This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Annealed and quenched representations of the Gauss-Rényi measure by “periodic points”

Shintaro Suzuki and Hiroki Takahasi Department of Mathematics, Tokyo Gakugei University, 4-1-1 Nukuikita-machi Koganei-shi, Tokyo, 184-8501, JAPAN shin05@u-gakugei.ac.jp Keio Institute of Pure and Applied Sciences (KiPAS), Department of Mathematics, Keio University, Yokohama, 223-8522, JAPAN hiroki@math.keio.ac.jp
Abstract.

We consider independently identically distributed random compositions of the Gauss and Rényi maps that generate random continued fractions. Using methods of ergodic theory, thermodynamic formalism and large deviations, we show that weighted cycles of this random dynamical system equidistribute with respect to the Gauss-Rényi measure. We present both annealed (sample-averaged) and quenched (samplewise) results.

2020 Mathematics Subject Classification:
11K50, 37A40, 37A44, 37C40
Keywords: random dynamical system; periodic points; the Gauss map; the Rényi map

1. Introduction

One leading idea in the qualitative theory of deterministic dynamical systems is to use the collection of periodic orbits as a spine to structure the dynamics. This idea traces back to Poincaré [32]: “… ce qui nous rend ces solutions périodiques si précieuses, … la seul brèche par où nous puissions esseyer de pénétrer dans une place jusqu’ici réputée inabordable.” Bowen’s pioneering results [7, 8] assert that periodic points of topologically mixing Axiom A diffeomorphisms equidistribute with respect to the measure of maximal entropy. The importance of periodic orbits in descriptions of ergodic properties of natural invariant probability measures has long been recognized in the physics literature, see e.g., [10, 17]. Cvitanović [10] proposed expansions of dynamical characteristics into series or products that consist of infinitely many periodic orbits, to better analyze the characteristics taking advantage of the simple structure of each periodic orbit in the expansions.

By deterministic dynamical systems, we mean ordinary differential equations or iterated maps. Systems with multiple evolution laws, called random dynamical systems [5], are also relevant to consider. For a large class of random dynamical systems, we expect that periodic orbits still play significant roles, but it is not clear how periodic points should be defined.

In discrete time, deterministic dynamical systems are iterations of one fixed map, whereas random dynamical systems are compositions of different maps chosen at random. A naive idea is to use fixed points of random compositions of nn maps as substitutes for periodic points of period nn. Such “periodic points” have been indeed considered, see e.g., [9, 33, 37]. For other substitutes for the concept of periodic points in the context random dynamical systems, see e.g., [13, 21, 25].

In [37], the authors proved an analogue of Bowen’s equidistribution theorem [7, 8] for random dynamical systems generated by a class of interval maps with finitely many branches. The aim of this paper is to extend this analogue to random dynamical systems generated by the Gauss and Rényi maps. The Gauss map T0:(0,1][0,1)T_{0}\colon(0,1]\to[0,1) and the Rényi map T1:[0,1)[0,1)T_{1}\colon[0,1)\to[0,1) are respectively given by

T0x=1x1xandT1x=11x11x.T_{0}x=\frac{1}{x}-\left\lfloor\frac{1}{x}\right\rfloor\quad\text{and}\quad T_{1}x=\frac{1}{1-x}-\left\lfloor\frac{1}{1-x}\right\rfloor.

The graph of T1T_{1} is obtained by reversing the graph of T0T_{0} around the axis {x=1/2}\{x=1/2\}, as shown in Figure 1. Since both maps have infinitely many branches, the random dynamical systems they generate are beyond the scope of [37].

Refer to caption
Refer to caption
Figure 1. The graph of the Gauss map T0T_{0} (left) and that of the Rényi map T1T_{1} (right): T01(0)={1/k:k}T_{0}^{-1}(0)=\{1/k\colon k\in\mathbb{N}\}, T11(0)={(k1)/k:k}T_{1}^{-1}(0)=\{(k-1)/k\colon k\in\mathbb{N}\}; T01(1)=T11(1)=T_{0}^{-1}(1)=T_{1}^{-1}(1)=\emptyset; T10=0T_{1}0=0, T10=1T_{1}^{\prime}0=1.

For a sample path ω=(ωn)n=1\omega=(\omega_{n})_{n=1}^{\infty} in the product space Ω={0,1}\Omega=\{0,1\}^{\mathbb{N}} of the discrete space {0,1}\{0,1\}, we consider a random composition

Tωn=TωnTωn1Tω1 for n.T_{\omega}^{n}=T_{\omega_{n}}\circ T_{\omega_{n-1}}\circ\cdots\circ T_{\omega_{1}}\ \text{ for }n\in\mathbb{N}.

Write Tω0T_{\omega}^{0} for the identity map on [0,1][0,1]. Let Λω\Lambda_{\omega} denote the set of x[0,1]x\in[0,1] such that TωnxT_{\omega}^{n}x is defined for every nn\in\mathbb{N}. Each xΛωx\in\Lambda_{\omega} has a continued fraction expansion

(1.1) x=ω1+(1)ω1  C1(ω,x)+(1)ω2  C2(ω,x)+(1)ω3  C3(ω,x)+,x=\omega_{1}+\frac{\displaystyle{\hfill{(-1)^{\omega_{1}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{1}(\omega,x)}\hfill}}+\frac{\displaystyle{\hfill{(-1)^{\omega_{2}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{2}(\omega,x)}\hfill}}+\frac{\displaystyle{\hfill{(-1)^{\omega_{3}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{3}(\omega,x)}\hfill}}+\cdots,

where each Cn(ω,x)C_{n}(\omega,x), nn\in\mathbb{N} is a positive integer that is determined by Tωn1xT_{\omega}^{n-1}x, ωn\omega_{n}, ωn+1\omega_{n+1}, and satisfies (1)ωn+1+Cn(ω,x)1(-1)^{\omega_{n+1}}+C_{n}(\omega,x)\geq 1 (see §\S2.1 for details). This type of continued fractions was first considered by Perron [29]. In the case ωn=0\omega_{n}=0 for all nn\in\mathbb{N} we obtain the well-known regular continued fraction

x=1  A1(x)+1  A2(x)+1  A3(x)+,x=\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{A_{1}(x)}\hfill}}+\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{A_{2}(x)}\hfill}}+\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{A_{3}(x)}\hfill}}+\cdots,

where An(x)=1/T0n1xA_{n}(x)=\lfloor 1/T_{0}^{n-1}x\rfloor for nn\in\mathbb{N}. In the case ωn=1\omega_{n}=1 for all nn\in\mathbb{N} we obtain the backward continued fraction

x=11  B1(x)1  B2(x)1  B3(x),x=1-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{B_{1}(x)}\hfill}}-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{B_{2}(x)}\hfill}}-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{B_{3}(x)}\hfill}}-\cdots,

where Bn(x)=1/(1T1n1x)+1B_{n}(x)=\lfloor 1/(1-T_{1}^{n-1}x)\rfloor+1 for nn\in\mathbb{N}. The backward continued fraction was used, for example, in computing certain inhomogeneous approximation constants [31]. For its connection with geodesic flows, see [3].

It is the essential difference between statistical properties of the sequences (An(x))n=1(A_{n}(x))_{n=1}^{\infty} and (Bn(x))n=1(B_{n}(x))_{n=1}^{\infty} that makes the random continued fraction interesting. For Lebesgue almost every irrational xx in (0,1)(0,1), each positive integer kk appears in (An(x))n=1(A_{n}(x))_{n=1}^{\infty} with frequency 1log2log(k+1)2k(k+2)\frac{1}{\log 2}\log\frac{(k+1)^{2}}{k(k+2)}, while the frequency of 22 in (Bn(x))n=1(B_{n}(x))_{n=1}^{\infty} is 11. This is due to the fact that T0T_{0} leaves invariant the Gauss measure dλ0=1log2dxx+1d\lambda_{0}=\frac{1}{\log 2}\frac{dx}{x+1}, while T1T_{1} leaves invariant the infinite measure dxx\frac{dx}{x}. More precisely, x=0x=0 is a neutral fixed point of T1T_{1}: T10=0T_{1}0=0 and T10=1T_{1}^{\prime}0=1. For more comparisons of the regular and backward continued fractions as well as more information on the singular behavior of the digit sequence in the backward continued fraction, see [1, 2, 19, 38, 42] for example.

1.1. Statements of results

We consider an independently identically distributed (i.i.d.) random dynamical system generated by T0T_{0} and T1T_{1}. This means that T1T_{1} is chosen with a fixed probability p(0,1)p\in(0,1) at each step. Let mpm_{p} denote the Bernoulli measure on the sample space Ω\Omega associated with the probability vector (1p,p)(1-p,p). By [18, Theorem 5.2], there exists a unique Borel probability measure λp\lambda_{p} on [0,1][0,1] that is absolutely continuous with respect to the Lebesgue measure on [0,1][0,1] and satisfies μ=(1p)μT01+pμT11\mu=(1-p)\cdot\mu\circ T_{0}^{-1}+p\cdot\mu\circ T_{1}^{-1}. The measure λp\lambda_{p}, called the Gauss-Rényi measure, is significant since for mpm_{p}-almost every ωΩ\omega\in\Omega and Lebesgue almost every xΛωx\in\Lambda_{\omega}, we have

limn1ni=0n1f(Tωix)=f𝑑λp for any continuous f:[0,1].\lim_{n\to\infty}\frac{1}{n}\sum_{i=0}^{n-1}f(T_{\omega}^{i}x)=\int fd\lambda_{p}\ \text{ for any continuous $f\colon[0,1]\to\mathbb{R}$.}

For p[0,1)p\in[0,1), let hp:[0,1][0,)h_{p}\colon[0,1]\to[0,\infty) denote the Radon-Nikodým derivative of λp\lambda_{p} with respect to the Lebesgue measure on [0,1][0,1]. We know that h0(x)=1log21x+1h_{0}(x)=\frac{1}{\log 2}\frac{1}{x+1}. For any p(0,1)p\in(0,1), hph_{p} is bounded from above and away from 0 [23, Proposition 3.4]. An explicit formula for hph_{p} is desired, since it is related to the frequency of digits in the random continued fraction expansion (2.1). Up to present, no algebraic formula for hph_{p} is known except for the case p=0p=0. Kalle et al. proved that hph_{p} is CC^{\infty} for any p(0,1)p\in(0,1) [24]. Bousoun et al. [6] obtained a functional-analytic formula for hph_{p} for p(0,1)p\in(0,1) sufficiently near 0.

Our aim here is to represent λp\lambda_{p} and hph_{p} for any p(0,1)p\in(0,1), using the collection of “periodic points”

ωΩn=1Fix(Tωn),Fix(Tωn)={xΛω:Tωnx=x}.\bigcup_{\omega\in\Omega}\bigcup_{n=1}^{\infty}{\rm Fix}(T_{\omega}^{n}),\ \ {\rm Fix}(T_{\omega}^{n})=\{x\in\Lambda_{\omega}\colon T_{\omega}^{n}x=x\}.

Elements of this set are called random cycles [37]. We first present a quenched (samplewise) representation, and then an annealed (sample-averaged) one. For ωΩ\omega\in\Omega and nn\in\mathbb{N} define

(1.2) Zω,n=xFix(Tωn)|(Tωn)x|1,Z_{\omega,n}=\sum_{x\in{\rm Fix}(T_{\omega}^{n})}\!\!\!\!\!\!|(T_{\omega}^{n})^{\prime}x|^{-1},

which plays the role of a normalizing constant. The derivatives of T0T_{0} and T1T_{1} at their discontinuities are the one-sided derivatives. For a topological space XX, let (X)\mathcal{M}(X) denote the space of Borel probability measures on XX endowed with the weak* topology. For ωΩ\omega\in\Omega, xΛωx\in\Lambda_{\omega} and nn\in\mathbb{N}, let Vnω(x)([0,1])V_{n}^{\omega}(x)\in\mathcal{M}([0,1]) denote the uniform probability distribution on the random orbit (Tωix)i=0n1(T_{\omega}^{i}x)_{i=0}^{n-1}. For p{0,1}p\in\{0,1\}, let mpm_{p} denote the Borel probability measure on Ω\Omega that is the unit point mass at the point p=pppp^{\infty}=ppp\cdots in Ω\Omega. Let λ1([0,1])\lambda_{1}\in\mathcal{M}([0,1]) denote the unit point mass at 0.

Theorem 1.1 (quenched representation of the Gauss-Rényi measure).

Let p(0,1)p\in(0,1). The following statements hold:

  • (a)

    for mpm_{p}-almost every ωΩ\omega\in\Omega and any continuous function F:([0,1])F\colon\mathcal{M}([0,1])\to\mathbb{R},

    limn1Zω,nxFix(Tωn)|(Tωn)x|1F(Vnω(x))=F(λp);\lim_{n\to\infty}\frac{1}{Z_{\omega,n}}\sum_{x\in{\rm Fix}(T_{\omega}^{n})}|(T^{n}_{\omega})^{\prime}x|^{-1}F(V_{n}^{\omega}(x))=F(\lambda_{p});
  • (b)

    for mpm_{p}-almost every ωΩ\omega\in\Omega and any continuous function f:[0,1]f\colon[0,1]\to\mathbb{R},

    limn1Zω,nxFix(Tωn)|(Tωn)x|1f𝑑Vnω(x)=f𝑑λp.\lim_{n\to\infty}\frac{1}{Z_{\omega,n}}\sum_{x\in{\rm Fix}(T^{n}_{\omega})}|(T_{\omega}^{n})^{\prime}x|^{-1}\int fdV_{n}^{\omega}(x)=\int fd\lambda_{p}.

As already noted, the cases p=0p=0 and p=1p=1 correspond to the iteration of T0T_{0} and that of T1T_{1} respectively. The convergences in Theorem 1.1 in these two cases were established in [40] (see [15] for a closely related result) and [42] respectively. The main concern of this paper is the case p(0,1)p\in(0,1).

Theorem 1.1(a) implies Theorem 1.1(b) (see §2.4). The latter deserves to be called a quenched representation of λp\lambda_{p} in terms of random cycles. For ωΩ\omega\in\Omega, xΛωx\in\Lambda_{\omega}, a subset AA of [0,1][0,1] and nn\in\mathbb{N}, let

en(ω,x,A)=#{0in1:TωixA}n.e_{n}(\omega,x,A)=\frac{\#\{0\leq i\leq n-1\colon T_{\omega}^{i}x\in A\}}{n}.

By the portmanteau theorem, Theorem 1.1(b) is equivalent to the following: for mpm_{p}-almost every ωΩ\omega\in\Omega and any Borel subset AA of [0,1][0,1] with λp(A)=0\lambda_{p}(\partial A)=0,

(1.3) limn1Zω,nxFix(Tωn)|(Tωn)x|1en(ω,x,A)=λp(A).\lim_{n\to\infty}\frac{1}{Z_{\omega,n}}\sum_{x\in{\rm Fix}(T^{n}_{\omega})}|(T_{\omega}^{n})^{\prime}x|^{-1}e_{n}(\omega,x,A)=\lambda_{p}(A).

The meaning of Theorem 1.1(a) may be a little less intuitive Theorem 1.1(b). By the portmanteau theorem it is equivalent to the following: for for mpm_{p}-almost every ωΩ\omega\in\Omega and any Borel subset 𝒜\mathcal{A} of (Λ)\mathcal{M}(\Lambda) with λp𝒜\lambda_{p}\notin\partial\mathcal{A},

limn1Zω,nxFix(Tωn)Vnω(x)𝒜|(Tωn)x|1=1l𝒜(λp),\lim_{n\to\infty}\frac{1}{Z_{\omega,n}}\sum_{\begin{subarray}{c}x\in{\rm Fix}(T_{\omega}^{n})\\ V_{n}^{\omega}(x)\in\mathcal{A}\end{subarray}}|(T^{n}_{\omega})^{\prime}x|^{-1}=\mbox{1}\hskip-2.5pt\mbox{l}_{\mathcal{A}}(\lambda_{p}),

where 1l𝒜\mbox{1}\hskip-2.5pt\mbox{l}_{\mathcal{A}} denotes the indicator function of 𝒜\mathcal{A}. In particular, if λp𝒜\lambda_{p}\in\mathcal{A} then Vnω(x)𝒜V_{n}^{\omega}(x)\in\mathcal{A} holds for almost every xFix(Tωn)x\in{\rm Fix}(T_{\omega}^{n}) as nn\to\infty.

To move on to an annealed counterpart, for p[0,1]p\in[0,1], nn\in\mathbb{N} and ωΩ\omega\in\Omega we set

Zp,n=Zω,n𝑑mp(ω),Z_{p,n}=\int Z_{\omega,n}dm_{p}(\omega),

which plays the role of a normalizing constant.

Theorem 1.2 (annealed representation of the Gauss-Rényi measure).

Let p(0,1)p\in(0,1). The following statements hold:

  • (a)

    for any continuous function F:([0,1])F\colon\mathcal{M}([0,1])\to\mathbb{R},

    limn1Zp,n𝑑mp(ω)xFix(Tωn)|(Tωn)x|1F(Vnω(x))=F(λp);\lim_{n\to\infty}\frac{1}{Z_{p,n}}\int dm_{p}(\omega)\sum_{x\in{\rm Fix}(T_{\omega}^{n})}|(T_{\omega}^{n})^{\prime}x|^{-1}F(V_{n}^{\omega}(x))=F(\lambda_{p});
  • (b)

    for any continuous function f:[0,1]f\colon[0,1]\to\mathbb{R},

    limn1Zp,n𝑑mp(ω)xFix(Tωn)|(Tωn)x|1f𝑑Vnω(x)=f𝑑λp.\lim_{n\to\infty}\frac{1}{Z_{p,n}}\int dm_{p}(\omega)\sum_{x\in{\rm Fix}(T_{\omega}^{n})}|(T_{\omega}^{n})^{\prime}x|^{-1}\int fdV^{\omega}_{n}(x)=\int fd\lambda_{p}.

Theorem 1.2(a) implies Theorem 1.2(b) (see §2.3). The latter deserves to be called an annealed representation of λp\lambda_{p} in terms of random cycles since it is equivalent to the following: for any Borel subset AA of [0,1][0,1] with λp(A)=0\lambda_{p}(\partial A)=0,

(1.4) limn1Zp,n𝑑mp(ω)xFix(Tωn)|(Tωn)x|1en(ω,x,A)=λp(A).\lim_{n\to\infty}\frac{1}{Z_{p,n}}\int dm_{p}(\omega)\sum_{x\in{\rm Fix}(T^{n}_{\omega})}|(T_{\omega}^{n})^{\prime}x|^{-1}e_{n}(\omega,x,A)=\lambda_{p}(A).

Theorem 1.2(a) is equivalent to the following: for any Borel subset 𝒜\mathcal{A} of (Λ)\mathcal{M}(\Lambda) with λp𝒜\lambda_{p}\notin\partial\mathcal{A},

limn1Zp,n𝑑mp(ω)xFix(Tωn)Vnω(x)𝒜|(Tωn)x|1=1l𝒜(λp).\lim_{n\to\infty}\frac{1}{Z_{p,n}}\int dm_{p}(\omega)\sum_{\begin{subarray}{c}x\in{\rm Fix}(T_{\omega}^{n})\\ V_{n}^{\omega}(x)\in\mathcal{A}\end{subarray}}|(T^{n}_{\omega})^{\prime}x|^{-1}=\mbox{1}\hskip-2.5pt\mbox{l}_{\mathcal{A}}(\lambda_{p}).

Since the Radon-Nikodým derivative hph_{p} of the Gauss-Rényi measure λp\lambda_{p} is continuous, from (1.3) and (1.4) we obtain its quenched and annealed representations in terms of random cycles.

Corollary 1.3 (quenched and annealed representations of the Radon-Nikodým derivative).

Let p(0,1)p\in(0,1). The following statements hold:

  • (a)

    for mpm_{p}-almost every ωΩ\omega\in\Omega and any y(0,1)y\in(0,1),

    hp(y)=limε+012εlimn1Zω,nxFix(Tωn)|(Tωn)x|1en(ω,x,[yε,y+ε]);h_{p}(y)=\lim_{\varepsilon\to+0}\frac{1}{2\varepsilon}\lim_{n\to\infty}\frac{1}{Z_{\omega,n}}\sum_{x\in{\rm Fix}(T^{n}_{\omega})}|(T_{\omega}^{n})^{\prime}x|^{-1}e_{n}(\omega,x,[y-\varepsilon,y+\varepsilon]);
  • (b)

    for any y(0,1)y\in(0,1),

    hp(y)=limε+012εlimn1Zp,n𝑑mp(ω)xFix(Tωn)|(Tωn)x|1en(ω,x,[yε,y+ε]).h_{p}(y)=\lim_{\varepsilon\to+0}\frac{1}{2\varepsilon}\lim_{n\to\infty}\frac{1}{Z_{p,n}}\int dm_{p}(\omega)\sum_{x\in{\rm Fix}(T_{\omega}^{n})}|(T_{\omega}^{n})^{\prime}x|^{-1}e_{n}(\omega,x,[y-\varepsilon,y+\varepsilon]).

Our main results altogether assert that the collection of random cycles capture relevant information of the Gauss-Rényi random dynamics. Since random cycles can be defined for general random dynamical systems, their relevance in descriptions of random dynamical properties should be investigated in a much more broader context. Our main results support the relevance, while Buzzi [9] earlier proved that a dynamical zeta function defined with random cycles of certain random matrices cannot be extended beyond its disk of holomorphy, almost surely. Under suitable assumptions, dynamical zeta functions of deterministic dynamical systems can be extended to meromorphic functions, and their zeros/poles are related to statistical properties of the underlying dynamics. With our results including [37] and Buzzi’s one [9] in mind, which information is captured by random cycles and which is not should be closely examined in the future.

1.2. Method of proofs of the main results

A basic strategy for proofs of our main results is to represent the i.i.d. random dynamical system generated by T0T_{0} and T1T_{1} as a skew product, and analyze the corresponding deterministic dynamical system. Let θ:ΩΩ\theta\colon\Omega\to\Omega denote the left shift: (θω)n=ωn+1(\theta\omega)_{n}=\omega_{n+1} for nn\in\mathbb{N}. Let

E={(ω,x)Ω×[0,1]:(ω1,x){(0,0),(1,1)}},E=\{(\omega,x)\in\Omega\times[0,1]\colon(\omega_{1},x)\in\{(0,0),(1,1)\}\},

and define R:(Ω×[0,1])EΩ×[0,1]R\colon(\Omega\times[0,1])\setminus E\to\Omega\times[0,1] by

R(ω,x)=(θω,Tω1x).R(\omega,x)=(\theta\omega,T_{\omega_{1}}x).

Let

Λ=n=0Rn((Ω×[0,1])E),\Lambda=\bigcap_{n=0}^{\infty}R^{-n}\left((\Omega\times[0,1])\setminus E\right),

which is a non-compact set. We still denote R|ΛR|_{\Lambda} by RR and call it the Gauss-Rényi map. We have Rn(ω,x)=(θnω,Tωnx)R^{n}(\omega,x)=(\theta^{n}\omega,T_{\omega}^{n}x) for (ω,x)Λ(\omega,x)\in\Lambda and nn\in\mathbb{N}, and so

Λω={x[0,1]:(ω,x)Λ}\Lambda_{\omega}=\{x\in[0,1]\colon(\omega,x)\in\Lambda\}

for every ωΩ\omega\in\Omega. For any p[0,1]p\in[0,1], the map RR leaves invariant the Borel probability measure mpλpm_{p}\otimes\lambda_{p}, the restriction of the product measure of mpm_{p} and λp\lambda_{p} to Λ\Lambda.

For each nn\in\mathbb{N}, let Fix(Rn){\rm Fix}(R^{n}) denote the set of periodic points of RR of period nn. A key observation is that xFix(Tωn)x\in{\rm Fix}(T_{\omega}^{n}) implies (ω,x)Fix(Rn)(\omega^{\prime},x)\in{\rm Fix}(R^{n}) where ωΩ\omega^{\prime}\in\Omega is the repetition of the word ω1ωn\omega_{1}\cdots\omega_{n} in ω\omega. For this reason, properties of random cycles may be analyzed through the analysis of periodic points of RR. Much of our effort is devoted to establishing annealed and quenched level-2 large deviations upper bounds for periodic points of RR, and derive the desired convergences from the large deviations upper bounds. For p[0,1]p\in[0,1], nn\in\mathbb{N} and ωΩ\omega\in\Omega, define

Qpn(ω)=(1p)#{1in:ωi=0}p#{1in:ωi=1},Q_{p}^{n}(\omega)=(1-p)^{\#\{1\leq i\leq n\colon\omega_{i}=0\}}p^{\#\{1\leq i\leq n\colon\omega_{i}=1\}},

where we put 00=10^{0}=1 for convenience. Notice that

(1.5) Zp,n=(ω,x)Fix(Rn)Qpn(ω)|(Tωn)x|1.Z_{p,n}=\sum_{(\omega,x)\in{\rm Fix}(R^{n})}Q_{p}^{n}(\omega)|(T^{n}_{\omega})^{\prime}x|^{-1}.

For (ω,x)Λ(\omega,x)\in\Lambda and nn\in\mathbb{N}, let VnR(ω,x)(Λ)V_{n}^{R}(\omega,x)\in\mathcal{M}(\Lambda) denote the uniform probability distribution on the orbit (Ri(ω,x))i=0n1(R^{i}(\omega,x))_{i=0}^{n-1}. Let δVnR(ω,x)\delta_{V_{n}^{R}(\omega,x)} denote the Borel probability measure on (Λ)\mathcal{M}(\Lambda) that is the unit point mass at VnR(ω,x)V_{n}^{R}(\omega,x). Define a sequence (μ~n)n=1(\tilde{\mu}_{n})_{n=1}^{\infty} of Borel probability measures on (Λ)\mathcal{M}(\Lambda) by

μ~n=1Zp,n(ω,x)Fix(Rn)Qpn(ω)|(Tωn)x|1δVnR(ω,x).\tilde{\mu}_{n}=\frac{1}{Z_{p,n}}\sum_{(\omega,x)\in{\rm Fix}(R^{n})}Q_{p}^{n}(\omega)|(T^{n}_{\omega})^{\prime}x|^{-1}\delta_{V_{n}^{R}(\omega,x)}.
Theorem 1.4 (annealed level-2 Large Deviation Principle).

Let p(0,1)p\in(0,1). The following statements hold:

  • (a)

    (μ~n)n=1(\tilde{\mu}_{n})_{n=1}^{\infty} is exponentially tight, and satisfies the LDP with the convex good rate function Ip:(Λ)[0,]:I_{p}\colon\mathcal{M}(\Lambda)\to[0,\infty]: for any open subset 𝒢\mathcal{G} of (Λ)\mathcal{M}(\Lambda),

    lim infn1nlogμ~n(𝒢)inf𝒢Ip,\liminf_{n\to\infty}\frac{1}{n}\log\tilde{\mu}_{n}(\mathcal{G})\geq-\inf_{\mathcal{G}}I_{p},

    and for any closed subset 𝒞\mathcal{C} of (Λ)\mathcal{M}(\Lambda),

    lim supn1nlogμ~n(𝒞)inf𝒞Ip.\limsup_{n\to\infty}\frac{1}{n}\log\tilde{\mu}_{n}(\mathcal{C})\leq-\inf_{\mathcal{C}}I_{p}.

    The minimizer of IpI_{p} is unique and it is mpλpm_{p}\otimes\lambda_{p};

  • (b)

    for any bounded continuous function F:(Λ)F\colon\mathcal{M}(\Lambda)\to\mathbb{R},

    limn1Zp,n(ω,x)Fix(Rn)Qpn(ω)|(Tωn)x|1F(VnR(ω,x))=F(mpλp).\lim_{n\to\infty}\frac{1}{Z_{p,n}}\sum_{(\omega,x)\in{\rm Fix}(R^{n})}Q_{p}^{n}(\omega)|(T^{n}_{\omega})^{\prime}x|^{-1}F(V_{n}^{R}(\omega,x))=F(m_{p}\otimes\lambda_{p}).

See §\S2.2 for the definition of the Large Deviation Principle and that of related terms in the statements of Theorem 1.4, including the meaning of level-2. The statements in the cases p=0p=0 and p=1p=1 were established in [40] and [42] respectively. The main concern of this paper is the case p(0,1)p\in(0,1).

Moving on to a quenched counterpart, for each ωΩ\omega\in\Omega we define a sequence (μ~nω)n=1(\tilde{\mu}_{n}^{\omega})_{n=1}^{\infty} of Borel probability measures on (Λ)\mathcal{M}(\Lambda) by

μ~nω=1Zω,nxFix(Tωn)|(Tωn)x|1δVnR(ω,x).\tilde{\mu}_{n}^{\omega}=\frac{1}{Z_{\omega,n}}\sum_{x\in{\rm Fix}(T^{n}_{\omega})}|(T_{\omega}^{n})^{\prime}x|^{-1}\delta_{V_{n}^{R}(\omega,x)}.

The measure Ωμ~nω()𝑑mp(ω)\int_{\Omega}\tilde{\mu}^{\omega}_{n}(\cdot)dm_{p}(\omega) on (Λ)\mathcal{M}(\Lambda) equals μ~n()\tilde{\mu}_{n}(\cdot) up to subexponential factors (see Lemma 3.7).

Theorem 1.5 (quenched level-2 large deviations).

Let p(0,1)p\in(0,1). The following statements hold:

  • (a)

    for mpm_{p}-almost every ωΩ\omega\in\Omega, (μ~nω)n=1(\tilde{\mu}_{n}^{\omega})_{n=1}^{\infty} is exponentially tight, and for any closed subset 𝒞\mathcal{C} of (Λ)\mathcal{M}(\Lambda),

    lim supn1nlogμ~nω(𝒞)inf𝒞Ip;\limsup_{n\to\infty}\frac{1}{n}\log\tilde{\mu}_{n}^{\omega}(\mathcal{C})\leq-\inf_{\mathcal{C}}I_{p};
  • (b)

    for mpm_{p}-almost every ωΩ\omega\in\Omega and any bounded continuous function F:(Λ)F\colon\mathcal{M}(\Lambda)\to\mathbb{R},

    limn1Zω,nxFix(Tωn)|(Tωn)x|1F(VnR(ω,x))=F(mpλp).\lim_{n\to\infty}\frac{1}{Z_{\omega,n}}\sum_{x\in{\rm Fix}(T^{n}_{\omega})}|(T_{\omega}^{n})^{\prime}x|^{-1}F(V_{n}^{R}(\omega,x))=F(m_{p}\otimes\lambda_{p}).

The rest of this paper consists of three sections. In §2 we prove Theorem 1.1 and Theorem 1.2 subject to Theorem 1.4 and Theorem 1.5. These deductions are rather straightforward. In §3 we start an analysis of the Gauss-Rényi map RR, and prove Theorem 1.5 subject to Theorem 1.4. In §4 we prove Theorem 1.4.

A more precise logical structure is indicated in the diagram below. In §2.3 we show Theorem 1.4(b) \Longrightarrow Theorem 1.2. In §2.4 we show Theorem 1.5(b) \Longrightarrow Theorem 1.1. In §3.5 we show Theorem 1.4(a) \Longrightarrow Theorem 1.5(a) \Longrightarrow Theorem 1.5(b).

𝐓𝐡𝐞𝐨𝐫𝐞𝐦1.4(𝐚)§3.5𝐓𝐡𝐞𝐨𝐫𝐞𝐦1.5(𝐚)§4.7§3.5𝐓𝐡𝐞𝐨𝐫𝐞𝐦1.4(𝐛)𝐓𝐡𝐞𝐨𝐫𝐞𝐦1.5(𝐛)§2.3§2.4𝐓𝐡𝐞𝐨𝐫𝐞𝐦1.2𝐓𝐡𝐞𝐨𝐫𝐞𝐦1.1{\begin{CD}{\bf Theorem~\ref{level-2-thm}(a)}@>{\S\ref{pf-sample}}>{}>{\bf Theorem~\ref{ldpup-q}(a)}\\ @V{\S\ref{section4-1}}V{}V@V{}V{\S\ref{pf-sample}}V\\ {\bf Theorem~\ref{level-2-thm}(b)}{\bf Theorem~\ref{ldpup-q}(b)}\\ @V{\S\ref{pf-thmb}}V{}V@V{}V{\S\ref{pf-thma}}V\\ {\bf Theorem~\ref{thm-b}}{\bf Theorem~\ref{thm-a}}\\ \end{CD}}

Most of our effort is dedicated to the proof of Theorem 1.4(a). The random dynamical system we consider falls into the class of mean expanding systems that are comprehensively investigated in [4]. Moreover, the restriction of the Perron-Frobenius operator associated with the Gauss-Rényi map RR to an appropriate function space has a spectral gap [23, 24]. This property can be used to apply the general results in [4] to deduce nice statistical properties of the dynamical system (Λ,R,mpλp)(\Lambda,R,m_{p}\otimes\lambda_{p}), see [23] for details. Meanwhile, it is not known whether the existence of spectral gap implies the LDP. To prove Theorem 1.4(a), our strategy is to code the Gauss-Rényi map into the countable full shift, establish the LDP there, and then transfer this LDP back to the original system.

Owing to the existence of the neutral fixed point of the Rényi map T1T_{1}, for the potential function associated with this countable full shift there exists no Gibbs state. To resolve this difficulty, we construct an appropriate induced system that is topologically conjugate to another countable full shift, and then apply the result of the second-named author in [42]. This requires verifying the regularity of the associated induced potential.

The uniqueness of minimizer in Theorem 1.4(a) is important to ensure the convergence in Theorem 1.4(b). To establish this uniqueness, we first show the uniqueness of equilibrium state (see Proposition 4.14), and then show that any minimizer is an equilibrium state. The first step relies on implementing the thermodynamic formalism for countable Markov shifts (see e.g., [27, 34]) with the induced system. Except for the construction of induced system and the verification of regularity of induced potential, the argument follows well-known lines (see e.g., [27, 30]). In the second step we appeal to the result of the second named author [40].

2. Deduction of convergences on random cycles

As a warm up, in §\S2.1 we begin by describing an induction algorithm that generates random continued fractions. In §\S2.2 we summarize basic facts on large deviations. We show Theorem 1.4(b) \Longrightarrow Theorem 1.2 and Theorem 1.5(b) \Longrightarrow Theorem 1.1, respectively in §2.3 and §2.4. Those readers who would like to immediately access the proofs of Theorems 1.1 and 1.2 can pass §2.1, §2.2 and directly go to §2.3 and §2.4.

Notation. For a bounded interval JJ, let |J||J| denote its Euclidean length.

2.1. A continued fraction algorithm by the Gauss-Rényi map

Using the Gauss-Rényi map, we describe an induction algorithm generating random continued fractions. Define a function C:(Ω×[0,1])EC\colon(\Omega\times[0,1])\setminus E\to\mathbb{N} by

C(ω,x)=1(1)ω1x+ω1.C(\omega,x)=\left\lfloor\frac{1}{(-1)^{\omega_{1}}x+\omega_{1}}\right\rfloor.

For (ω,x)(Ω×[0,1])E(\omega,x)\in(\Omega\times[0,1])\setminus E and nn\in\mathbb{N}, let

Cn(ω,x)=C(Rn1(ω,x))+ωn+1,C_{n}(\omega,x)=C(R^{n-1}(\omega,x))+\omega_{n+1},

when Rn1(ω,x)R^{n-1}(\omega,x) is defined.

For any (ω,x)(Ω×[0,1])E(\omega,x)\in(\Omega\times[0,1])\setminus E we have

x=ω1+(1)ω1C(ω,x)+Tω1x.x=\omega_{1}+\frac{(-1)^{\omega_{1}}}{C(\omega,x)+T_{\omega_{1}}x}.

If R(ω,x)ER(\omega,x)\notin E, then replacing (ω,x)(\omega,x) in (2.1) by R(ω,x)R(\omega,x) we have

Tω1x=ω2+(1)ω2C(R(ω,x))+Tω2x.T_{\omega_{1}}x=\omega_{2}+\frac{(-1)^{\omega_{2}}}{C(R(\omega,x))+T_{\omega}^{2}x}.

Substituting this into the right-hand side of the previous equality yields

x=ω1+(1)ω1  C(ω,x)+ω2+(1)ω2  C(R(ω,x))+Tω2x.x=\omega_{1}+\frac{\displaystyle{\hfill{(-1)^{\omega_{1}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C(\omega,x)+\omega_{2}}\hfill}}+\frac{\displaystyle{\hfill{(-1)^{\omega_{2}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C(R(\omega,x))+T_{\omega}^{2}x}\hfill}}.

If n2n\geq 2 and Ri(ω,x)ER^{i}(\omega,x)\notin E for i=0,,n1i=0,\ldots,n-1, then repeating the above process yields

x=ω1+(1)ω1  C1(ω,x)++(1)ωn1  Cn1(ω,x)+(1)ωn  Cn(ω,x)ωn+1+Tωnx,x=\omega_{1}+\frac{\displaystyle{\hfill{(-1)^{\omega_{1}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{1}(\omega,x)}\hfill}}+\cdots+\frac{\displaystyle{\hfill{(-1)^{\omega_{n-1}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{n-1}(\omega,x)}\hfill}}+\frac{\displaystyle{\hfill{(-1)^{\omega_{n}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{n}(\omega,x)-\omega_{n+1}+T_{\omega}^{n}x}\hfill}},

where (1)ωi+1+Ci(ω,x)1(-1)^{\omega_{i+1}}+C_{i}(\omega,x)\geq 1 for i=1,,ni=1,\ldots,n.

For many (ω,x)(\omega,x), this algorithm produces a continued fraction expansion of xx summarized as follows.

Proposition 2.1.

Let (ω,x)(Ω×[0,1])E(\omega,x)\in(\Omega\times[0,1])\setminus E.

  • (a)

    If xΛωx\in\Lambda_{\omega}, then (1)ωn+1+Cn(ω,x)1(-1)^{\omega_{n+1}}+C_{n}(\omega,x)\geq 1 for every nn\in\mathbb{N}, and the continued fraction

    ω1+(1)ω1  C1(ω,x)+(1)ω2  C2(ω,x)+(1)ω3  C3(ω,x)+\omega_{1}+\frac{\displaystyle{\hfill{(-1)^{\omega_{1}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{1}(\omega,x)}\hfill}}+\frac{\displaystyle{\hfill{(-1)^{\omega_{2}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{2}(\omega,x)}\hfill}}+\frac{\displaystyle{\hfill{(-1)^{\omega_{3}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{3}(\omega,x)}\hfill}}+\cdots

    converges to xx.

  • (b)

    If xΛωx\in\Lambda_{\omega}, then xx\notin\mathbb{Q} if and only if (1)ωn+1+Cn(ω,x)2(-1)^{\omega_{n+1}}+C_{n}(\omega,x)\geq 2 for infinitely many nn\in\mathbb{N}.

  • (c)

    If xΛωx\notin\Lambda_{\omega} then xx\in\mathbb{Q}.

To prove (a) and (b) we use the next lemma. For related results, see [26, 29, 43].

Lemma 2.2 ([28, Lemma 2.1(a)]).

Let ωΩ\omega\in\Omega and (Cn)n(C_{n})_{n\in\mathbb{N}}\in\mathbb{N}^{\mathbb{N}} satisfy (1)ωn+1+Cn1(-1)^{\omega_{n+1}}+C_{n}\geq 1 for every nn\in\mathbb{N}. Then the continued fraction

ω1+(1)ω1  C1+(1)ω2  C2+(1)ω3  C3+\omega_{1}+\frac{\displaystyle{\hfill{(-1)^{\omega_{1}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{1}}\hfill}}+\frac{\displaystyle{\hfill{(-1)^{\omega_{2}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{2}}\hfill}}+\frac{\displaystyle{\hfill{(-1)^{\omega_{3}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{3}}\hfill}}+\cdots

converges to a number in [0,1][0,1]. This number is irrational if and only if (1)ωn+1+Cn2(-1)^{\omega_{n+1}}+C_{n}\geq 2 for infinitely many nn\in\mathbb{N}.

Proof of Proposition 2.1.

Let xΛωx\in\Lambda_{\omega}. Applying the algorithm to (ω,x)(\omega,x) we get

(2.1) x=ω1+(1)ω1C(ω,x)+Tω1x,x=\omega_{1}+\frac{(-1)^{\omega_{1}}}{C(\omega,x)+T_{\omega_{1}}x},

and for every n2n\geq 2,

(2.2) x=ω1+(1)ω1  C1(ω,x)++(1)ωn1  Cn1(ω,x)+(1)ωn  Cn(ω,x)ωn+1+Tωnx,x=\omega_{1}+\frac{\displaystyle{\hfill{(-1)^{\omega_{1}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{1}(\omega,x)}\hfill}}+\cdots+\frac{\displaystyle{\hfill{(-1)^{\omega_{n-1}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{n-1}(\omega,x)}\hfill}}+\frac{\displaystyle{\hfill{(-1)^{\omega_{n}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{n}(\omega,x)-\omega_{n+1}+T_{\omega}^{n}x}\hfill}},

where (1)ωi+1+Ci(ω,x)1(-1)^{\omega_{i+1}}+C_{i}(\omega,x)\geq 1 for i=1,,ni=1,\ldots,n. By Lemma 2.2, the continued fraction

ω1+(1)ω1  C1(ω,x)+(1)ω2  C2(ω,x)+(1)ω3  C3(ω,x)+\omega_{1}+\frac{\displaystyle{\hfill{(-1)^{\omega_{1}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{1}(\omega,x)}\hfill}}+\frac{\displaystyle{\hfill{(-1)^{\omega_{2}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{2}(\omega,x)}\hfill}}+\frac{\displaystyle{\hfill{(-1)^{\omega_{3}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{3}(\omega,x)}\hfill}}+\cdots

converges to a number y[0,1]y\in[0,1]. For (a) and (b) it suffices to show x=yx=y.

For each nn\in\mathbb{N}, let Jn(ω,x)J_{n}(\omega,x) denote the maximal subinterval of [0,1][0,1] containing xx on which TωnT_{\omega}^{n} is monotone. From (2.2) we have yJn(ω,x)y\in J_{n}(\omega,x) for every nn\in\mathbb{N}. Since (1)ωn+1+Cn(ω,x)1(-1)^{\omega_{n+1}}+C_{n}(\omega,x)\geq 1, there are four cases:

  • (i)

    ωn=ωn+1=0\omega_{n}=\omega_{n+1}=0;

  • (ii)

    ωn=1\omega_{n}=1 and ωn+1=0\omega_{n+1}=0;

  • (iii)

    ωn=0\omega_{n}=0, C(Rn1(ω,x))2C(R^{n-1}(\omega,x))\geq 2 and ωn+1=1\omega_{n+1}=1;

  • (iv)

    ωn=ωn+1=1\omega_{n}=\omega_{n+1}=1.

We estimate the derivatives of the composition using the definitions of T0T_{0} and T1T_{1}, inf(0,1]|T0|1\inf_{(0,1]}|T^{\prime}_{0}|\geq 1 and inf[0,1)|T1|1\inf_{[0,1)}|T^{\prime}_{1}|\geq 1, the monotonicity of |T0||T_{0}| on (0,1](0,1] and that of |T1||T_{1}^{\prime}| on [0,1)[0,1). In case (i), for all yTωn1Jn(ω,x)y\in T_{\omega}^{n-1}J_{n}(\omega,x) we have

|(Tωn+1Tωn)y||T0(23)|=94.|(T_{\omega_{n+1}}\circ T_{\omega_{n}})^{\prime}y|\geq\left|T_{0}^{\prime}\left(\frac{2}{3}\right)\right|=\frac{9}{4}.

In case (ii), for all yTωn1Jn(ω,x)y\in T_{\omega}^{n-1}J_{n}(\omega,x) we have

|(Tωn+1Tωn)y||T1(13)|=94.|(T_{\omega_{n+1}}\circ T_{\omega_{n}})^{\prime}y|\geq\left|T_{1}^{\prime}\left(\frac{1}{3}\right)\right|=\frac{9}{4}.

In case (iii), for all yTωn1Jn(ω,x)y\in T_{\omega}^{n-1}J_{n}(\omega,x) we have

|(Tωn+1Tωn)y||T0(12)|>94.|(T_{\omega_{n+1}}\circ T_{\omega_{n}})^{\prime}y|\geq\left|T_{0}^{\prime}\left(\frac{1}{2}\right)\right|>\frac{9}{4}.

Hence, if one of (i) (ii) (iii) occurs infinitely many times then infJn(ω,x)|(Tωn)|\inf_{J_{n}(\omega,x)}|(T_{\omega}^{n})^{\prime}|\to\infty as nn\to\infty. By the mean value theorem, for every nn\in\mathbb{N} there exists ξnJn(ω,x)\xi_{n}\in J_{n}(\omega,x) such that

|xy|=|TωnxTωny||(Tωn)ξn|1|(Tωn)ξn|.|x-y|=\frac{|T_{\omega}^{n}x-T_{\omega}^{n}y|}{|(T_{\omega}^{n})^{\prime}\xi_{n}|}\leq\frac{1}{|(T_{\omega}^{n})^{\prime}\xi_{n}|}.

Letting nn\to\infty we obtain x=yx=y.

If all (i) (ii) (iii) occur only finitely many times, then there is kk\in\mathbb{N} such that ωn=1\omega_{n}=1 for every n>kn>k. Suppose TωkxT_{\omega}^{k}x\notin\mathbb{Q}. Then T1n(Tωkx)0T_{1}^{n}(T_{\omega}^{k}x)\neq 0 holds for every nn\in\mathbb{N}. Then the formula for T1T_{1} implies infJnk(1,Tωkx)|(T1nk)|\inf_{J_{n-k}(1^{\infty},T_{\omega}^{k}x)}|(T_{1}^{n-k})^{\prime}|\to\infty as nn\to\infty. For every nn\in\mathbb{N} there exists ζnJnk(1,Tωkx)\zeta_{n}\in J_{n-k}(1^{\infty},T_{\omega}^{k}x) such that

|TωkxTωky|=|TωnxTωny||(T1nk)ζn|1|(T1nk)ζn|.|T_{\omega}^{k}x-T_{\omega}^{k}y|=\frac{|T_{\omega}^{n}x-T_{\omega}^{n}y|}{|(T_{1}^{n-k})^{\prime}\zeta_{n}|}\leq\frac{1}{|(T_{1}^{n-k})^{\prime}\zeta_{n}|}.

Letting nn\to\infty we obtain Tωkx=TωkyT_{\omega}^{k}x=T_{\omega}^{k}y. Since the restriction of TωkT_{\omega}^{k} to Jk(ω,x)J_{k}(\omega,x) is injective, we obtain x=yx=y. Suppose TωkxT_{\omega}^{k}x\in\mathbb{Q}. Since T1T_{1} maps all rational points to 0, there exists nn\in\mathbb{N} such that T1n(Tωkx)=0T_{1}^{n}(T_{\omega}^{k}x)=0. Since the neutral fixed point 0 of T1T_{1} is topologically repelling, it follows that T1n(Tωky)=0T_{1}^{n}(T_{\omega}^{k}y)=0. The restriction of Tωk+nT_{\omega}^{k+n} to Jk+n(ω,x)J_{k+n}(\omega,x) is injective, and hence x=yx=y. We have verified (a) and (b).

If x(0,1)Λωx\in(0,1)\setminus\Lambda_{\omega} then there exists nn\in\mathbb{N} such that TωnxT^{n}_{\omega}x is defined and Tωn+1xT^{n+1}_{\omega}x is not defined. Then Tωnx{0,1}T^{n}_{\omega}x\in\{0,1\} holds and (2.1), (2.2) together imply xx\in\mathbb{Q}, verifying (c). The proof of Proposition 2.1 is complete. ∎

2.2. Large Deviation Principle

Our main reference on large deviations is [11]. Let 𝒳\mathcal{X} be a topological space and let (μn)n=1(\mu_{n})_{n=1}^{\infty} be a sequence of Borel probability measures on 𝒳\mathcal{X}. We say the Large Deviation Principle (LDP) holds for (μn)n=1(\mu_{n})_{n=1}^{\infty} if there exists a lower semicontinuous function I:𝒳[0,]I\colon\mathcal{X}\to[0,\infty] such that:

  • (a)

    for any open subset 𝒢\mathcal{G} of 𝒳\mathcal{X},

    lim infn1nlogμn(𝒢)inf𝒢I;\liminf_{n\to\infty}\frac{1}{n}\log\mu_{n}(\mathcal{G})\geq-\inf_{\mathcal{G}}I;
  • (b)

    for any closed subset 𝒞\mathcal{C} of 𝒳\mathcal{X},

    lim supn1nlogμn(𝒞)inf𝒞I.\limsup_{n\to\infty}\frac{1}{n}\log\mu_{n}(\mathcal{C})\leq-\inf_{\mathcal{C}}I.

We say x𝒳x\in\mathcal{X} is a minimizer if I(x)=0I(x)=0 holds. The LDP roughly means that in the limit nn\to\infty the measure μn\mu_{n} assigns all but exponentially small mass to the set {x𝒳:I(x)=0}\{x\in\mathcal{X}\colon I(x)=0\} of minimizers. The function II is called a rate function. If 𝒳\mathcal{X} is a metric space and (μn)n=1(\mu_{n})_{n=1}^{\infty} satisfies the LDP, the rate function is unique. We say the rate function II is good if the set {x𝒳:I(x)c}\{x\in\mathcal{X}\colon I(x)\leq c\} is compact for any c>0c>0.

We say (μn)n=1(\mu_{n})_{n=1}^{\infty} is exponentially tight if for any L>0L>0 there exists a compact subset 𝒦\mathcal{K} of 𝒳\mathcal{X} such that

lim supn1nlogμn(𝒳𝒦)L.\limsup_{n\to\infty}\frac{1}{n}\log\mu_{n}(\mathcal{X}\setminus\mathcal{K})\leq-L.

If (μn)n=1(\mu_{n})_{n=1}^{\infty} is exponentially tight then it is tight, i.e., for any ε>0\varepsilon>0 there exists a compact subset 𝒦\mathcal{K}^{\prime} of 𝒳\mathcal{X} such that μn(𝒦)>1ε\mu_{n}(\mathcal{K}^{\prime})>1-\varepsilon for all sufficiently large nn.

Proposition 2.3.

Let 𝒳\mathcal{X}, 𝒴\mathcal{Y} be Hausdorff spaces and let (μn)n=1(\mu_{n})_{n=1}^{\infty} be a sequence of Borel probability measures on 𝒳\mathcal{X} for which the LDP holds with a good rate function II. Let f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} be a continuous map. Then the LDP holds for (μnf1)n=1(\mu_{n}\circ f^{-1})_{n=1}^{\infty} with a good rate function J:𝒴[0,]J\colon\mathcal{Y}\to[0,\infty] given by

J(y)=inf{I(x):x𝒳,f(x)=y}.J(y)=\inf\{I(x)\colon x\in\mathcal{X},\ f(x)=y\}.

Moreover, if y0𝒴y_{0}\in\mathcal{Y} is a mininizer of JJ, then there is a minimizer x0𝒳x_{0}\in\mathcal{X} of II such that y0=f(x0)y_{0}=f(x_{0}).

The first assertion of Proposition 2.3 is well-known as the Contraction Principle. Here we only include a proof of the second assertion.

Proof of the second assertion of Proposition 2.3.

Let y0𝒴y_{0}\in\mathcal{Y} be a minimizer of JJ. By the definition of JJ, there is a sequence (xn)n=1(x_{n})_{n=1}^{\infty} in 𝒳\mathcal{X} such that y0=f(xn)y_{0}=f(x_{n}) and I(xn)<1/nI(x_{n})<1/n for every n1n\geq 1. Since II is a good rate function, (xn)n=1(x_{n})_{n=1}^{\infty} has a limit point, say x0x_{0}. Since II is lower semicontinuous, x0x_{0} is a minimizer of II. Since ff is continuous, we obtain y0=f(x0)y_{0}=f(x_{0}). ∎

Let XX be a topological space and let C(X)C(X) denote the Banach space of real-valued bounded continuous functions on XX endowed with the supremum norm. Recall that the weak* topology on (X)\mathcal{M}(X) is the coarsest topology that makes the map μ(X)f𝑑μ\mu\in\mathcal{M}(X)\mapsto\int fd\mu continuous for any fC(X)f\in C(X). In this topology, a sequence (μn)n=1(\mu_{n})_{n=1}^{\infty} of elements of (X)\mathcal{M}(X) converges to μ(X)\mu\in\mathcal{M}(X) if and only if limnf𝑑μn=f𝑑μ\lim_{n}\int fd\mu_{n}=\int fd\mu holds for any fC(X)f\in C(X). This condition is equivalent to limnf𝑑μn=f𝑑μ\lim_{n}\int fd\mu_{n}=\int fd\mu for any fC(X)f\in C(X) that is uniformly continuous (see [36, Chapter 9]).

Donsker and Varadhan have identified three levels of the LDP, see e.g., [12, Chapter I]. The LDP for a sequence of Borel probability measures on (X)\mathcal{M}(X) is referred to as level-2. The LDP for a sequence of Borel probability measures on \mathbb{R} determined by a real-valued function on XX is referred to as level-1. By the Contraction Principle, any level-2 LDP can be transferred to a level-1 LDP.

Notation. For a topological space XX, let 2(X)\mathcal{M}^{2}(X) denote the space of Borel probability measures on (X)\mathcal{M}(X) endowed with the weak* topology. For each μ(X)\mu\in\mathcal{M}(X), let δμ2(X)\delta_{\mu}\in\mathcal{M}^{2}(X) denote the unit point mass at μ\mu.

2.3. Proof of Theorem 1.2

We define a sequence (ξ~n)n=1(\tilde{\xi}_{n})_{n=1}^{\infty} in 2([0,1])\mathcal{M}^{2}([0,1]) by

ξ~n=1Zp,n𝑑mp(ω)xFix(Tωn)|(Tωn)x|1δVnω(x).\tilde{\xi}_{n}=\frac{1}{Z_{p,n}}\int dm_{p}(\omega)\sum_{x\in{\rm Fix}(T_{\omega}^{n})}|(T_{\omega}^{n})^{\prime}x|^{-1}\delta_{V_{n}^{\omega}(x)}.

Also, we define a sequence (ξn)n=1(\xi_{n})_{n=1}^{\infty} in ([0,1])\mathcal{M}([0,1]) by

ξn=1Zp,n𝑑mp(ω)xFix(Tωn)|(Tωn)x|1Vnω(x).\xi_{n}=\frac{1}{Z_{p,n}}\int dm_{p}(\omega)\sum_{x\in{\rm Fix}(T_{\omega}^{n})}|(T_{\omega}^{n})^{\prime}x|^{-1}V^{\omega}_{n}(x).

The convergence in Theorem 1.2(a) is equivalent to the convergence of (ξ~n)n=1(\tilde{\xi}_{n})_{n=1}^{\infty} to δλp\delta_{\lambda_{p}} in 2(Λ)\mathcal{M}^{2}(\Lambda). The convergence in Theorem 1.2(b) is equivalent to the convergence of (ξn)n=1(\xi_{n})_{n=1}^{\infty} to λp\lambda_{p} in ([0,1])\mathcal{M}([0,1]).

Let Π:Ω×[0,1][0,1]\Pi\colon\Omega\times[0,1]\to[0,1] be the projection to the second coordinate. The restriction of Π\Pi to Λ\Lambda induces a continuous map Π:μ(Λ)μΠ1([0,1])\Pi_{*}\colon\mu\in\mathcal{M}(\Lambda)\mapsto\mu\circ\Pi^{-1}\in\mathcal{M}([0,1]), which induces a continuous map μ~2(Λ)μ~Π12([0,1])\tilde{\mu}\in\mathcal{M}^{2}(\Lambda)\mapsto\tilde{\mu}\circ\Pi_{*}^{-1}\in\mathcal{M}^{2}([0,1]). Note that Π(μ)=ν\Pi_{*}(\mu)=\nu implies δμΠ1=δν.\delta_{\mu}\circ\Pi_{*}^{-1}=\delta_{\nu}. In particular, δmpλpΠ1=δλp\delta_{m_{p}\otimes\lambda_{p}}\circ\Pi_{*}^{-1}=\delta_{\lambda_{p}} and δVnR((ω,x))Π1=δVnω(x)\delta_{V_{n}^{R}((\omega,x))}\circ\Pi_{*}^{-1}=\delta_{V_{n}^{\omega}(x)} for (ω,x)Fix(Rn)(\omega,x)\in{\rm Fix}(R^{n}), and the latter yields μ~nΠ1=ξ~n.\tilde{\mu}_{n}\circ\Pi_{*}^{-1}=\tilde{\xi}_{n}. By Theorem 1.4(b), (μ~n)n=1(\tilde{\mu}_{n})_{n=1}^{\infty} converges to δmpλp\delta_{m_{p}\otimes\lambda_{p}} in 2(Λ)\mathcal{M}^{2}(\Lambda), and hence (ξ~n)n=1(\tilde{\xi}_{n})_{n=1}^{\infty} converges to δλp\delta_{\lambda_{p}} in 2([0,1])\mathcal{M}^{2}([0,1]) as required in Theorem 1.2(a).

We define a continuous map Ξ:2([0,1])([0,1])\Xi\colon\mathcal{M}^{2}([0,1])\to\mathcal{M}([0,1]) as follows. For each μ~2([0,1])\tilde{\mu}\in\mathcal{M}^{2}([0,1]), consider the positive normalized bounded linear functional on C([0,1])C([0,1]) given by

fC([0,1])(f𝑑μ)𝑑μ~(μ).f\in C([0,1])\mapsto\int\left(\int fd\mu\right)d\tilde{\mu}(\mu).

Using Riesz’s representation theorem, we define Ξ(μ~)\Xi(\tilde{\mu}) to be the unique element of ([0,1])\mathcal{M}([0,1]) such that

f𝑑Ξ(μ~)=(f𝑑μ)𝑑μ~(μ) for all fC([0,1]).\int fd\Xi(\tilde{\mu})=\int\left(\int fd\mu\right)d\tilde{\mu}(\mu)\ \text{ for all $f\in C([0,1])$.}

Clearly Ξ\Xi is continuous, satisfies Ξ(ξ~n)=ξn\Xi(\tilde{\xi}_{n})=\xi_{n} for every nn\in\mathbb{N} and Ξ(δλp)=λp\Xi(\delta_{\lambda_{p}})=\lambda_{p}. Hence, Theorem 1.2(b) follows from Theorem 1.2(a). ∎

2.4. Proof of Theorem 1.1

For each ωΩ\omega\in\Omega, define a sequence (ξnω)n=1({\xi}^{\omega}_{n})_{n=1}^{\infty} in 2([0,1])\mathcal{M}^{2}([0,1]) by

ξ~nω=1Zω,nxFix(Tωn)|(Tωn)x|1δVnω(x).\tilde{\xi}^{\omega}_{n}=\frac{1}{Z_{\omega,n}}\sum_{x\in{\rm Fix}(T^{n}_{\omega})}|(T_{\omega}^{n})^{\prime}x|^{-1}\delta_{V_{n}^{\omega}(x)}.

Also, define a sequence (ξnω)n=1({\xi}^{\omega}_{n})_{n=1}^{\infty} in ([0,1])\mathcal{M}([0,1]) by

ξnω=1Zω,nxFix(Tωn)|(Tωn)x|1Vnω(x).{\xi}^{\omega}_{n}=\frac{1}{Z_{\omega,n}}\sum_{x\in{\rm Fix}(T^{n}_{\omega})}|(T_{\omega}^{n})^{\prime}x|^{-1}V_{n}^{\omega}(x).

The convergence in Theorem 1.1(a) is equivalent to the convergence of (ξ~nω)n=1(\tilde{\xi}^{\omega}_{n})_{n=1}^{\infty} to δλp\delta_{\lambda_{p}} in 2([0,1])\mathcal{M}^{2}([0,1]). The convergence in Theorem 1.1(b) is equivalent to the convergence of (ξnω)n=1(\xi^{\omega}_{n})_{n=1}^{\infty} to λp\lambda_{p} in ([0,1])\mathcal{M}([0,1]).

To finish, we trace the proof of Theorem 1.2. By Theorem 1.5(b), (μ~nω)n=1(\tilde{\mu}_{n}^{\omega})_{n=1}^{\infty} converges to δmpλp\delta_{m_{p}\otimes\lambda_{p}} in 2(Λ)\mathcal{M}^{2}(\Lambda). Since μ~nωΠ1=ξ~nω\tilde{\mu}_{n}^{\omega}\circ\Pi_{*}^{-1}=\tilde{\xi}_{n}^{\omega}, (ξ~nω)n=1(\tilde{\xi}_{n}^{\omega})_{n=1}^{\infty} converges to δλp\delta_{\lambda_{p}} in 2([0,1])\mathcal{M}^{2}([0,1]) as required in Theorem 1.1(a). Since Ξ(ξ~nω)=ξnω\Xi(\tilde{\xi}_{n}^{\omega})=\xi_{n}^{\omega} and Ξ(δλp)=λp\Xi(\delta_{\lambda_{p}})=\lambda_{p}, (ξnω)n=1(\xi_{n}^{\omega})_{n=1}^{\infty} converges to λp\lambda_{p} in ([0,1])\mathcal{M}([0,1]) as required in Theorem 1.1(b). ∎

3. Fundamental analysis of the Gauss-Rényi map

In this section we start the analysis of the Gauss-Rényi map RR. In §3.1 we introduce an inducing scheme and some related objects. In §3.2 we introduce an induced map R^\widehat{R} and investigate its expansion properties. In §3.3 we introduce an annealed geometric potential φ\varphi and evaluate distortions of its Birkhoff averages. In §3.4 we prove several preliminary lemmas needed for the proof of Theorem 1.5. The proof of Theorem 1.5 is given in §3.5.

Convention. Since p(0,1)p\in(0,1) is a fixed constant for the rest of the paper, it will be mostly omitted from each statement.

3.1. Inducing scheme

An inducing scheme of a dynamical system T:XXT\colon X\to X is a pair (Y,tY)(Y,t_{Y}), where YY is a proper subset of XX and tY:Y{}t_{Y}\colon Y\to\mathbb{N}\cup\{\infty\} is a function given by

tY(x)=inf{n1:TnxY}.t_{Y}(x)=\inf\{n\geq 1\colon T^{n}x\in Y\}.

Given an inducing scheme (Y,tY)(Y,t_{Y}) of T:XXT\colon X\to X, for each kk\in\mathbb{N} we set

{tY=k}={xY:tY(x)=k},\{t_{Y}=k\}=\{x\in Y\colon t_{Y}(x)=k\},

and define an induced map

T^:k=1{tY=k}T^tY(x)xY,\widehat{T}\colon\bigcup_{k=1}^{\infty}\{t_{Y}=k\}\mapsto{\widehat{T}}^{t_{Y}(x)}x\in Y,

and define an inducing domain

X^=n=0T^n(k=1{tY=k}).\widehat{X}=\bigcap_{n=0}^{\infty}{\widehat{T}}^{-n}\left(\bigcup_{k=1}^{\infty}\{t_{Y}=k\}\right).

In other words, tYt_{Y} is the first return time to YY, T^\widehat{T} is the first return map to YY and X^\widehat{X} is the domain on which T^\widehat{T} can be iterated infinitely many times. We still denote by T^\widehat{T} the restriction of T^\widehat{T} to X^\widehat{X}. We call T^:X^X^\widehat{T}\colon\widehat{X}\to\widehat{X} an induced system associated with the inducing scheme (Y,tY)(Y,t_{Y}).

We will consider an induced system of the Gauss-Rényi map R:ΛΛR\colon\Lambda\to\Lambda and its symbolic version. We will attach the symbol “ ^\widehat{\cdot} ” to denote objects associated with inducing schemes.

3.2. Building uniform expansion

Let 0\mathbb{N}_{0} and 1\mathbb{N}_{1} denote the sets of even and odd positive integers respectively. A direct calculation shows that both T0T_{0} and T1T_{1} satisfy Rényi’s condition, namely

(3.1) sup(2k+2,2k]|T0′′||T0|22for all k0 and sup[k1k+1,k+1k+3)|T1′′||T1|22for all k1.\sup_{\left(\frac{2}{k+2},\frac{2}{k}\right]}\frac{|T_{0}^{\prime\prime}|}{|T_{0}^{\prime}|^{2}}\leq 2\quad\text{for all }k\in\mathbb{N}_{0}\ \text{ and }\ \sup_{\left[\frac{k-1}{k+1},\frac{k+1}{k+3}\right)}\frac{|T_{1}^{\prime\prime}|}{|T_{1}^{\prime}|^{2}}\leq 2\quad\text{for all }k\in\mathbb{N}_{1}.

Define a1:(Ω×[0,1])Ea_{1}\colon(\Omega\times[0,1])\setminus E\to\mathbb{N} by

(3.2) a1(ω,x)={k0 if ω1=0 and x(2k+2,2k],k1 if ω1=1 and x[k1k+1,k+1k+3).a_{1}(\omega,x)=\begin{cases}\vskip 2.84526ptk\in\mathbb{N}_{0}&\text{ if }\omega_{1}=0\text{ and }x\in\displaystyle{\left(\frac{2}{k+2},\frac{2}{k}\right]},\\ k\in\mathbb{N}_{1}&\text{ if }\omega_{1}=1\text{ and }x\in\displaystyle{\left[\frac{k-1}{k+1},\frac{k+1}{k+3}\right)}.\end{cases}

For each (ω,x)(\omega,x) and nn\in\mathbb{N} such that Rn1(ω,x)R^{n-1}(\omega,x) is defined, let

an(ω,x)=a1(Rn1(ω,x)).a_{n}(\omega,x)=a_{1}(R^{n-1}(\omega,x)).

For nn\in\mathbb{N} and a1anna_{1}\cdots a_{n}\in\mathbb{N}^{n}, define an nn-cylinder

Δ(a1an)={(ω,x)(Ω×[0,1])E:ai(ω,x)=ai for i=1,,n}.\varDelta(a_{1}\cdots a_{n})=\{(\omega,x)\in(\Omega\times[0,1])\setminus E\colon a_{i}(\omega,x)=a_{i}\text{ for }i=1,\ldots,n\}.

Let Π:Ω×[0,1][0,1]\Pi\colon\Omega\times[0,1]\to[0,1] denote the projection to the second coordinate. We write J(a1an)J(a_{1}\cdots a_{n}) for Π(Δ(a1an))\Pi(\varDelta(a_{1}\cdots a_{n})). If (ω,x)Δ(a1an)(\omega,x)\in\varDelta(a_{1}\cdots a_{n}) then J(a1an)J(a_{1}\cdots a_{n}) is the maximal subinterval of [0,1][0,1] containing xx on which TωnT_{\omega}^{n} is monotone. The collection of 11-cylinders defines a Markov partition for RR: for every kk\in\mathbb{N}, RR maps Δ(k)\varDelta(k) bijectively onto its image and R(Δ(k))R(\varDelta(k)) contains Ω×(0,1)\Omega\times(0,1).

Refer to caption
Figure 2. The inducing domain Λ^\widehat{\Lambda} associated with the inducing scheme (ΛΔ(1),tΛΔ(1))(\Lambda\setminus\varDelta(1),t_{\Lambda\setminus\varDelta(1)}) is contained in k=2Δ(k)\bigcup_{k=2}^{\infty}\varDelta(k), the shaded area.

Put

(3.3) Ω0={(ωn)nΩ:ωn=0 for infinitely many n}.\Omega_{0}=\{(\omega_{n})_{n\in\mathbb{N}}\in\Omega\colon\omega_{n}=0\text{ for infinitely many $n$}\}.

Due to the presence of the neutral fixed point of the Rényi map T1T_{1}, the random composition of T0T_{0} and T1T_{1} is not uniformly expanding in that

infωΩ0infΛωlim infn1nlog|(Tωn)|=0.\inf_{\omega\in\Omega_{0}}\inf_{\Lambda_{\omega}}\liminf_{n\to\infty}\frac{1}{n}\log|(T_{\omega}^{n})^{\prime}|=0.

To control the effect of the neutral fixed point, we consider the inducing scheme (ΛΔ(1),tΛΔ(1))(\Lambda\setminus\varDelta(1),t_{\Lambda\setminus\varDelta(1)}) of R:ΛΛR\colon\Lambda\to\Lambda and the associated induced system R^:Λ^Λ^\widehat{R}\colon\widehat{\Lambda}\to\widehat{\Lambda}, see Figure 2. Let us abbreviate tΛΔ(1)t_{\Lambda\setminus\varDelta(1)} as tt. Note that t(ω,x)t(\omega,x) is finite if and only if Tωx0T_{\omega}x\neq 0. The next lemma implies that the induced map R^\widehat{R} is still not uniformly expanding. However, the lemma after the next one implies that R^2\widehat{R}^{2} is uniformly expanding.

Lemma 3.1.

Let ωΩ\omega\in\Omega satisfy ω1=0\omega_{1}=0, ω2=1\omega_{2}=1, ω3=0\omega_{3}=0. Then we have

infxΔ(2)|(Tωt(ω,x))x|=1.\inf_{x\in\varDelta(2)}|(T_{\omega}^{t(\omega,x)})^{\prime}x|=1.
Proof.

Since inf(0,1]|T0|1\inf_{(0,1]}|T_{0}^{\prime}|\geq 1 and inf[0,1)|T1|1\inf_{[0,1)}|T_{1}^{\prime}|\geq 1, we have infxΔ(2)|(Tωt(ω,x))x|1\inf_{x\in\varDelta(2)}|(T_{\omega}^{t(\omega,x)})^{\prime}x|\geq 1. By the hypothesis on ω\omega and T01=0T_{0}1=0, we have limx10t(ω,x)=2\lim_{x\to 1-0}t(\omega,x)=2. Using this and the monotonicity of |T0||T_{0}^{\prime}| on Δ(2)\varDelta(2) and that of |T1||T_{1}^{\prime}| on Δ(1)\varDelta(1), we obtain infxΔ(2)|(Tωt(ω,x))x|limx10|(T1T0)x|=1\inf_{x\in\varDelta(2)}|(T_{\omega}^{t(\omega,x)})^{\prime}x|\leq\lim_{x\to 1-0}|(T_{1}\circ T_{0})^{\prime}x|=1. ∎

Lemma 3.2.

If (ω,x)ΛΔ(1)(\omega,x)\in\Lambda\setminus\varDelta(1), t(ω,x)t(\omega,x) and t(R^(ω,x))t(\widehat{R}(\omega,x)) are finite and ai(ω,x)=ai(ϱ,y)a_{i}(\omega,x)=a_{i}(\varrho,y) for i=1,,t(ω,x)+t(R^(ω,x))i=1,\ldots,t(\omega,x)+t(\widehat{R}(\omega,x)), then

|(Tωt(ω,x)+t(R^(ω,x)))y||(Tωt(ω,x)+t(R^(ω,x))1)(Tωy)|94.|(T_{\omega}^{t(\omega,x)+t(\widehat{R}(\omega,x))})^{\prime}y|\geq|(T_{\omega}^{t(\omega,x)+t(\widehat{R}(\omega,x))-1})^{\prime}(T_{\omega}y)|\geq\frac{9}{4}.
Proof.

From the definitions of T0T_{0} and T1T_{1}, inf(0,1]|T0|1\inf_{(0,1]}|T^{\prime}_{0}|\geq 1, inf[0,1)|T1|1\inf_{[0,1)}|T^{\prime}_{1}|\geq 1, the monotonicity of |T0||T_{0}| on (0,1](0,1] and that of |T1||T_{1}^{\prime}| on [0,1)[0,1), if (ω,x)Δ(2)(\omega,x)\notin\varDelta(2) then

(Tωt(ω,x)+t(R^(ω,x)))y||Tω1y||T0(12)|>94.(T_{\omega}^{t(\omega,x)+t(\widehat{R}(\omega,x))})^{\prime}y|\geq|T_{\omega_{1}}^{\prime}y|\geq\left|T_{0}^{\prime}\left(\frac{1}{2}\right)\right|>\frac{9}{4}.

If (ω,x)Δ(2)(\omega,x)\in\varDelta(2) and Tωt(ω,x)x[1/2,1)T_{\omega}^{t(\omega,x)}x\in[1/2,1) then

(Tωt(ω,x)+t(R^(ω,x)))y||Tt(ω,x)y||T1(13)|=94.(T_{\omega}^{t(\omega,x)+t(\widehat{R}(\omega,x))})^{\prime}y|\geq|T_{t(\omega,x)}^{\prime}y|\geq\left|T_{1}^{\prime}\left(\frac{1}{3}\right)\right|=\frac{9}{4}.

If (ω,x)Δ(2)(\omega,x)\in\varDelta(2) and Tωt(ω,x)x(0,1/2)T_{\omega}^{t(\omega,x)}x\in(0,1/2) then

(Tωt(ω,x)+t(R^(ω,x)))y||T(Tωt(ω,x)y)||T0(12)|>94.(T_{\omega}^{t(\omega,x)+t(\widehat{R}(\omega,x))})^{\prime}y|\geq|T^{\prime}(T_{\omega}^{t(\omega,x)}y)|\geq\left|T_{0}^{\prime}\left(\frac{1}{2}\right)\right|>\frac{9}{4}.

Hence the desired inequality holds. ∎

Lemma 3.3 (Uniform decay of cylinders).

There exists K1K\geq 1 such that for every nn\in\mathbb{N} and every a1anna_{1}\cdots a_{n}\in\mathbb{N}^{n},

|J(a1an)|Kn.|J(a_{1}\cdots a_{n})|\leq\frac{K}{\sqrt{n}}.
Proof.

Take an integer M4M\geq 4 such that for every nMn\geq M,

(3.4) (94)n/2+11n.\left(\frac{9}{4}\right)^{-\sqrt{n}/2+1}\leq\frac{1}{\sqrt{n}}.

Set K=M/2K=\sqrt{M}/2. Clearly we have |J(k)|1/2|J(k)|\leq 1/2 for every kk\in\mathbb{N}. Hence, for every 1nM1\leq n\leq M and every a1anna_{1}\cdots a_{n}\in\mathbb{N}^{n} we have |J(a1an)|1/2=K/MK/n|J(a_{1}\cdots a_{n})|\leq 1/2=K/\sqrt{M}\leq K/\sqrt{n} as required.

Let nM+1n\geq M+1 and a1anna_{1}\cdots a_{n}\in\mathbb{N}^{n}. We may assume a1ana_{1}\cdots a_{n} contains 11, for otherwise a direct calculation shows |J(a1an)|1/(n+1)|J(a_{1}\cdots a_{n})|\leq 1/(n+1). Let N1N\geq 1 denote the total number of blocks of consecutive 11s in a1ana_{1}\cdots a_{n}. A block of length not exceeding n\sqrt{n} is called a short block. A block which is not short is called a long block. If Nn/2N\geq\sqrt{n}/2, then Lemma 3.2 implies |J(a1an)|(9/4)n/2+1|J(a_{1}\cdots a_{n})|\leq(9/4)^{-\sqrt{n}/2+1}. This and (3.4) together yield the desired inequality.

Suppose N<n/2N<\sqrt{n}/2. If there is no long block, then #{1in:ai1}nnN>n/2\#\{1\leq i\leq n\colon a_{i}\neq 1\}\geq n-\sqrt{n}N>n/2. Let j=min{i1:ai1}j=\min\{i\geq 1\colon a_{i}\neq 1\} and k=max{i1:ai1}k=\max\{i\geq 1\colon a_{i}\neq 1\}. Define (ωi)iΩ(\omega_{i})_{i\in\mathbb{N}}\in\Omega by ωiaimod2\omega_{i}\equiv a_{i}\mod 2. By the mean value theorem and Lemma 3.2, for some 1\ell\geq 1 and all xTωj1(J(a1an))x\in T_{\omega}^{j-1}(J(a_{1}\cdots a_{n})) we have

1|Tθjωkj+1Tωj1(J(a1an))|=|Tθjωt(θjω,x)+t(R^(θjω,x))++t(R^1(θjω,x))Tωj1(J(a1an))|(94)/2|Tωj1(J(a1an))|(94)/2|J(a1an)|.\begin{split}1\geq&|T_{\theta^{j}\omega}^{k-j+1}\circ T_{\omega}^{j-1}(J(a_{1}\cdots a_{n}))|\\ =&|T_{\theta^{j}\omega}^{t(\theta^{j}\omega,x)+t(\widehat{R}(\theta^{j}\omega,x))+\cdots+t(\widehat{R}^{\ell-1}(\theta^{j}\omega,x))}\circ T_{\omega}^{j-1}(J(a_{1}\cdots a_{n}))|\\ \geq&\left(\frac{9}{4}\right)^{\lfloor\ell/2\rfloor}|T_{\omega}^{j-1}(J(a_{1}\cdots a_{n}))|\geq\left(\frac{9}{4}\right)^{\lfloor\ell/2\rfloor}|J(a_{1}\cdots a_{n})|.\end{split}

Since n/21n/22\ell\geq\lfloor n/2\rfloor-1\geq n/2-2 we have /2n/41\ell/2\geq n/4-1, and so /2n/41=n/41\lfloor\ell/2\rfloor\geq\lfloor n/4-1\rfloor=\lfloor n/4\rfloor-1. Combining this inequality with the above yields |J(a1an)|(9/4)n/4+1|J(a_{1}\cdots a_{n})|\leq(9/4)^{-\lfloor n/4\rfloor+1}. By nM+15n\geq M+1\geq 5 and (3.4), we obtain (9/4)n/4+1(9/4)n/2+11/n(9/4)^{-\lfloor n/4\rfloor+1}\leq(9/4)^{-\sqrt{n}/2+1}\leq 1/\sqrt{n}. If there is a long block, then there exists 1jn11\leq j\leq n-1 such that ai=1a_{i}=1 for i=j,,j+n1i=j,\ldots,j+\lfloor\sqrt{n}\rfloor-1, and thus Tωj1(J(a1an))J(1n)[0,1/(n+1)T_{\omega}^{j-1}(J(a_{1}\cdots a_{n}))\subset J(1^{\lfloor\sqrt{n}\rfloor})\subset[0,1/(\lfloor\sqrt{n}\rfloor+1). By the mean value theorem we obtain |J(a1an)|1/n|J(a_{1}\cdots a_{n})|\leq 1/\sqrt{n}. ∎

3.3. Annealed geometric potential

We introduce a function φ:(Ω×[0,1])E\varphi\colon(\Omega\times[0,1])\setminus E\to\mathbb{R} by

φ(ω,x)=logp(ω1)log|Tω1x|,\varphi(\omega,x)=\log p(\omega_{1})-\log|T_{\omega_{1}}^{\prime}x|,

where

p(ω1)={1p if ω1=0,p if ω1=1.p(\omega_{1})=\begin{cases}1-p\ &\text{ if }\omega_{1}=0,\\ p\ &\text{ if }\omega_{1}=1.\end{cases}

Note that φ\varphi is unbounded and supφ<0.\sup\varphi<0. We call φ\varphi an annealed geometric potential. For nn\in\mathbb{N} write SnφS_{n}\varphi for the Birkhoff sum i=0n1φRi\sum_{i=0}^{n-1}\varphi\circ R^{i}, and put S0φ0S_{0}\varphi\equiv 0 for convenience. The annealed geometric potential ties in with Theorem 1.2. For all (ω,x)Λ(\omega,x)\in\Lambda and all nn\in\mathbb{N} we have

exp(Snφ(ω,x))=Qnp(ω)|(Tωn)x|1.\exp(S_{n}\varphi(\omega,x))=Q_{n}^{p}(\omega)|(T_{\omega}^{n})^{\prime}x|^{-1}.

Compare this formula with (1.5). The next distortion estimate is straight forward.

Lemma 3.4.

For all nn\in\mathbb{N}, a1anna_{1}\cdots a_{n}\in\mathbb{N}^{n} and any pair (ω,x),(ϱ,y)(\omega,x),(\varrho,y) of points in Δ(a1an)\varDelta(a_{1}\cdots a_{n}),

Snφ(ω,x)Snφ(ϱ,y)2i=1n|TωixTϱiy|.S_{n}\varphi(\omega,x)-S_{n}\varphi(\varrho,y)\leq 2\sum_{i=1}^{n}|T^{i}_{\omega}x-T^{i}_{\varrho}y|.
Proof.

We have

Snφ(ω,x)Snφ(ϱ,y)=log|(Tωn)y||(Tωn)x|=log|(Tϱn)y||(Tϱn)x|.S_{n}\varphi(\omega,x)-S_{n}\varphi(\varrho,y)=\log\frac{|(T_{\omega}^{n})^{\prime}y|}{|(T_{\omega}^{n})^{\prime}x|}=\log\frac{|(T_{\varrho}^{n})^{\prime}y|}{|(T_{\varrho}^{n})^{\prime}x|}.

Then the desired inequality follows from the chain rule and (3.1). ∎

For each nn\in\mathbb{N} define

Dn(φ)=sup{Snφ(ω,x)Snφ(ϱ,y):ai(ω,x)=ai(ϱ,y),i=1,,n}.D_{n}(\varphi)=\sup\{S_{n}\varphi(\omega,x)-S_{n}\varphi(\varrho,y)\colon a_{i}(\omega,x)=a_{i}(\varrho,y),\ i=1,\ldots,n\}.

Note that D1(φ)<D_{1}(\varphi)<\infty, and Dn(φ)D_{n}(\varphi) is decreasing in nn.

Lemma 3.5.

We have Dn(φ)=O(n)D_{n}(\varphi)=O(\sqrt{n}) (n).(n\to\infty).

Proof.

Let nn\in\mathbb{N}, a1anna_{1}\cdots a_{n}\in\mathbb{N}^{n} and let (ω,x),(ϱ,y)Δ(a1an)(\omega,x),(\varrho,y)\in\varDelta(a_{1}\cdots a_{n}). Using Lemma 3.4 and then Lemma 3.3, we have

Snφ(ω,x)Snφ(φ,y)2i=1n|TωixTϱiy|2+2i=1n1|J(ai+1an)|Ki=1n1ni+1=O(n),\begin{split}S_{n}\varphi(\omega,x)-S_{n}\varphi(\varphi,y)&\leq 2\sum_{i=1}^{n}|T_{\omega}^{i}x-T_{\varrho}^{i}y|\\ &\leq 2+2\sum_{i=1}^{n-1}|J(a_{i+1}\cdots a_{n})|\leq K\sum_{i=1}^{n}\frac{1}{\sqrt{n-i+1}}=O(\sqrt{n}),\end{split}

which implies the assertion of the lemma. ∎

3.4. Preliminary lemmas for the proof of Theorem 1.5

One key point in the proof of Theorem 1.5 is that the measure Ωμ~nω()𝑑mp(ω)\int_{\Omega}\tilde{\mu}^{\omega}_{n}(\cdot)dm_{p}(\omega) equals μ~n()\tilde{\mu}_{n}(\cdot) up to subexponential factors. To show this, we first provide subexponential bounds on the normalizing constants Zω,nZ_{\omega,n} in (1.2).

Lemma 3.6.

For all ωΩ\omega\in\Omega and nn\in\mathbb{N} we have

exp(Dn(φ))Zω,nexp(Dn(φ)).\exp(-D_{n}(\varphi))\leq Z_{\omega,n}\leq\exp(D_{n}(\varphi)).

In particular, Zp,nZ_{p,n} is finite for all p(0,1)p\in(0,1) and all nn\in\mathbb{N}.

Proof.

Let ωΩ\omega\in\Omega, nn\in\mathbb{N} and let a1ana_{1}\cdots a_{n}\in\mathbb{N}^{\mathbb{N}} satisfy ωiai\omega_{i}\equiv a_{i} mod 22 for i=1,,ni=1,\ldots,n. Clearly, J(a1an)Fix(Tωn)J(a_{1}\cdots a_{n})\cap{\rm Fix}(T_{\omega}^{n}) is a singleton. Let x(a1an)x(a_{1}\cdots a_{n}) denote the element of this singleton. By the mean value theorem, for each a1anna_{1}\cdots a_{n}\in\mathbb{N}^{n} there exists y(a1an)J(a1an)y(a_{1}\cdots a_{n})\in J(a_{1}\cdots a_{n}) such that |(Tωn)y(a1an)|1=|J(a1an)|.|(T_{\omega}^{n})^{\prime}y(a_{1}\cdots a_{n})|^{-1}=|J(a_{1}\cdots a_{n})|. We have

exp(Dn(φ))|J(a1an)||(Tωn)x(a1an)|1exp(Dn(φ))|J(a1an)|.\exp(-D_{n}(\varphi))|J(a_{1}\cdots a_{n})|\leq|(T_{\omega}^{n})^{\prime}x(a_{1}\cdots a_{n})|^{-1}\leq\exp(D_{n}(\varphi))|J(a_{1}\cdots a_{n})|.

Summing the first inequality over all relevant a1ana_{1}\cdots a_{n} gives

Zω,nexp(Dn(φ))a1annaiωimod2i=1,,n|J(a1an)|=exp(Dn(φ)),Z_{\omega,n}\geq\exp(-D_{n}(\varphi))\sum_{\begin{subarray}{c}a_{1}\cdots a_{n}\in\mathbb{N}^{n}\\ a_{i}\equiv\omega_{i}\mod 2\\ \ i=1,\ldots,n\end{subarray}}|J(a_{1}\cdots a_{n})|=\exp(-D_{n}(\varphi)),

as required. Summing the second inequality in the double inequalities over all relevant a1ana_{1}\cdots a_{n} yields the required upper bound. ∎

Lemma 3.7.

For any Borel subset 𝒞\mathcal{C} of (Λ)\mathcal{M}(\Lambda) and every nn\in\mathbb{N},

exp(2Dn(φ))μ~n(𝒞)Ωμ~nω(𝒞)𝑑mp(ω)exp(2Dn(φ))μ~n(𝒞).\exp(-2D_{n}(\varphi))\tilde{\mu}_{n}(\mathcal{C})\leq\int_{\Omega}\tilde{\mu}_{n}^{\omega}(\mathcal{C})dm_{p}(\omega)\leq\exp(2D_{n}(\varphi))\tilde{\mu}_{n}(\mathcal{C}).
Proof.

By Lemma 3.6, for all ωΩ\omega\in\Omega and all nn\in\mathbb{N} we have

(3.5) exp(2Dn(φ))Zω,n/ΩZω,n𝑑mp(ω)exp(2Dn(φ)).\exp(-2D_{n}(\varphi))\leq{Z_{\omega,n}\Big{/}\int_{\Omega}Z_{\omega^{\prime},n}dm_{p}(\omega^{\prime})}\leq\exp(2D_{n}(\varphi)).

By the definitions of μ~n\tilde{\mu}_{n} and μ~nω\tilde{\mu}_{n}^{\omega}, for any Borel subset 𝒞\mathcal{C} of (Λ)\mathcal{M}(\Lambda) and all nn\in\mathbb{N},

(3.6) μ~n(𝒞)=1Zp,n(ω,x)Fix(Rn)VnR(ω,x)𝒞Qpn(ω)|(Tωn)x|1=ΩxFix(Tωn)VnR(ω,x)𝒞|(Tωn)x|1dmp(ω)/ΩZω,n𝑑mp(ω)=Ωμ~nω(𝒞)(Zω,n/ΩZω,n𝑑mp(ω))𝑑mp(ω).\begin{split}\tilde{\mu}_{n}(\mathcal{C})&=\frac{1}{Z_{p,n}}\sum_{\begin{subarray}{c}(\omega,x)\in{\rm Fix}(R^{n})\\ V_{n}^{R}(\omega,x)\in\mathcal{C}\end{subarray}}Q_{p}^{n}(\omega)|(T_{\omega}^{n})^{\prime}x|^{-1}\\ &=\int_{\Omega}\sum_{\begin{subarray}{c}x\in{\rm Fix}(T^{n}_{\omega})\\ V_{n}^{R}(\omega,x)\in\mathcal{C}\end{subarray}}|(T_{\omega}^{n})^{\prime}x|^{-1}dm_{p}(\omega)\Big{/}\int_{\Omega}Z_{\omega^{\prime},n}dm_{p}(\omega^{\prime})\\ &=\int_{\Omega}\tilde{\mu}_{n}^{\omega}(\mathcal{C})\left({Z_{\omega,n}\Big{/}\int_{\Omega}Z_{\omega^{\prime},n}dm_{p}(\omega^{\prime})}\right)dm_{p}(\omega).\end{split}

Combining (3.5) and (3.6) yields the desired inequality. ∎

The next lemma gives an upper bound for each closed subset of (Λ)\mathcal{M}(\Lambda) by the rate function IpI_{p}, but is not sufficient for Theorem 1.5(a) since the set of permissible samples depends on the closed set in consideration.

Lemma 3.8.

For any closed subset 𝒞\mathcal{C} of (Λ)\mathcal{M}(\Lambda), there exists a Borel subset Γ(𝒞)\Gamma(\mathcal{C}) of Ω\Omega such that mp(Γ(𝒞))=1m_{p}(\Gamma(\mathcal{C}))=1 and for every ωΓ(𝒞)\omega\in\Gamma(\mathcal{C}),

lim supn1nlogμ~nω(𝒞)inf𝒞Ip.\limsup_{n\to\infty}\frac{1}{n}\log\tilde{\mu}_{n}^{\omega}(\mathcal{C})\leq-\inf_{\mathcal{C}}I_{p}.
Proof.

Let 𝒞\mathcal{C} be a closed subset of (Λ)\mathcal{M}(\Lambda). We may assume inf𝒞Ip>0\inf_{\mathcal{C}}I_{p}>0, for otherwise the inequality is obvious. We first consider the case inf𝒞Ip<\inf_{\mathcal{C}}I_{p}<\infty. For ε(0,1)\varepsilon\in(0,1) and n1n\geq 1, set

Ωε,n={ωΩ:μ~nω(𝒞)exp(n(1ε)inf𝒞Ip)}.\Omega_{\varepsilon,n}=\left\{\omega\in\Omega\colon\tilde{\mu}_{n}^{\omega}(\mathcal{C})\geq\exp\left(-n(1-\varepsilon)\inf_{\mathcal{C}}I_{p}\right)\right\}.

By Markov’s inequality and the second inequality in Lemma 3.7,

mp(Ωε,n)exp(n(1ε)inf𝒞Ip)Ωμ~nω(𝒞)𝑑mp(ω)exp(2Dn(φ))exp(n(1ε)inf𝒞Ip)μ~n(𝒞).\begin{split}m_{p}(\Omega_{\varepsilon,n})&\leq\exp\left(n(1-\varepsilon)\inf_{\mathcal{C}}I_{p}\right)\int_{\Omega}\tilde{\mu}_{n}^{\omega}(\mathcal{C})dm_{p}(\omega)\\ &\leq\exp(2D_{n}(\varphi))\exp\left(n(1-\varepsilon)\inf_{\mathcal{C}}I_{p}\right)\tilde{\mu}_{n}(\mathcal{C}).\end{split}

By the LDP in Theorem 1.4(a), mp(Ωε,n)m_{p}(\Omega_{\varepsilon,n}) decays exponentially as nn increases. By Borel-Cantelli’s lemma, the inequality μ~nω(𝒞)exp(n(1ε)inf𝒞Ip)\tilde{\mu}_{n}^{\omega}(\mathcal{C})\geq\exp(-n(1-\varepsilon)\inf_{\mathcal{C}}I_{p}) holds only for finitely many nn for mpm_{p}-almost every ωΩ.\omega\in\Omega. Since ε(0,1)\varepsilon\in(0,1) is arbitrary, we obtain the desired inequality for mpm_{p}-almost every ωΩ\omega\in\Omega.

To treat the remaining case inf𝒞Ip=\inf_{\mathcal{C}}I_{p}=\infty, for k,nk,n\in\mathbb{N} we set

Ωk,n={ωΩ:μ~nω(𝒞)ekn}.\Omega_{k,n}=\left\{\omega\in\Omega\colon\tilde{\mu}_{n}^{\omega}(\mathcal{C})\geq e^{-kn}\right\}.

By Markov’s inequality and Lemma 3.7,

mp(Ωk,n)eknΩμ~nω(𝒞)𝑑mp(ω)exp(2Dn(φ))eknμ~n(𝒞).m_{p}(\Omega_{k,n})\leq e^{kn}\int_{\Omega}\tilde{\mu}_{n}^{\omega}(\mathcal{C})dm_{p}(\omega)\leq\exp(2D_{n}(\varphi))e^{kn}\tilde{\mu}_{n}(\mathcal{C}).

Since 𝒞\mathcal{C} is closed, the LDP in Theorem 1.4(a) gives lim supn(1/n)logμ~n(𝒞)inf𝒞Ip=\limsup_{n}(1/n)\log\tilde{\mu}_{n}(\mathcal{C})\leq-\inf_{\mathcal{C}}I_{p}=-\infty. Hence mp(Ωk,n)m_{p}(\Omega_{k,n}) decays exponentially as nn increases. By Borel-Cantelli’s lemma, there exists a Borel subset Γk(𝒞)\Gamma_{k}(\mathcal{C}) of Ω\Omega such that mp(Γk(𝒞))=1m_{p}(\Gamma_{k}(\mathcal{C}))=1, and for any ωΓk(𝒞)\omega\in\Gamma_{k}(\mathcal{C}) the inequality μ~nω(𝒞)ekn\tilde{\mu}_{n}^{\omega}(\mathcal{C})\geq e^{-kn} holds only for finitely many nn. Put Γ(𝒞)=k=1Γk(𝒞)\Gamma(\mathcal{C})=\bigcap_{k=1}^{\infty}\Gamma_{k}(\mathcal{C}). We have mp(Γ(𝒞))=1m_{p}(\Gamma(\mathcal{C}))=1, and lim supn(1/n)logμ~nω(𝒞)==inf𝒞Ip\limsup_{n}(1/n)\log\tilde{\mu}_{n}^{\omega}(\mathcal{C})=-\infty=-\inf_{\mathcal{C}}I_{p} for all ωΓ(𝒞)\omega\in\Gamma(\mathcal{C}) as required. ∎

Since (Λ)\mathcal{M}(\Lambda) is non-compact, we need the following auxiliary lemma that leads to the exponential tightness of (μ~nω)n=1(\tilde{\mu}^{\omega}_{n})_{n=1}^{\infty} as in Proposition 1.5(a).

Lemma 3.9.

For any L>0L>0 there exists a compact subset 𝒦L\mathcal{K}_{L} of (Λ)\mathcal{M}(\Lambda) and a Borel subset ΓL\Gamma_{L} of Ω\Omega such that mp(ΓL)=1m_{p}(\Gamma_{L})=1 and for every ωΓL\omega\in\Gamma_{L},

lim supn1nlogμ~nω((Λ)𝒦L)L.\limsup_{n\to\infty}\frac{1}{n}\log\tilde{\mu}^{\omega}_{n}(\mathcal{M}(\Lambda)\setminus\mathcal{K}_{L})\leq-L.
Proof.

By the exponential tightness of (μ~n)n=1(\tilde{\mu}_{n})_{n=1}^{\infty} in Theorem 1.4(a), for any L>0L>0 there is a compact subset 𝒦L\mathcal{K}_{L} of (Λ)\mathcal{M}(\Lambda) such that

(3.7) lim supn1nlogμ~n((Λ)𝒦L)2L.\limsup_{n\to\infty}\frac{1}{n}\log\tilde{\mu}_{n}(\mathcal{M}(\Lambda)\setminus\mathcal{K}_{L})\leq-2L.

For nn\in\mathbb{N}, set

ΩL,n={ωΩ:μ~nω((Λ)𝒦L)eLn}.\Omega_{L,n}=\left\{\omega\in\Omega\colon\tilde{\mu}_{n}^{\omega}(\mathcal{M}(\Lambda)\setminus\mathcal{K}_{L})\geq e^{-Ln}\right\}.

By Markov’s inequality and Lemma 3.7,

mp(ΩL,n)eLnΩμ~nω((Λ)𝒦L)𝑑mp(ω)exp(2Dn(φ))eLnμ~n((Λ)𝒦L).m_{p}(\Omega_{L,n})\leq e^{Ln}\int_{\Omega}\tilde{\mu}_{n}^{\omega}(\mathcal{M}(\Lambda)\setminus\mathcal{K}_{L})dm_{p}(\omega)\\ \leq\exp(2D_{n}(\varphi))e^{Ln}\tilde{\mu}_{n}(\mathcal{M}(\Lambda)\setminus\mathcal{K}_{L}).

By Lemma 3.5 and (3.7), mp(ΩL,n)m_{p}(\Omega_{L,n}) decays exponentially as nn increases. By Borel-Cantelli’s lemma, the number of those nn\in\mathbb{N} with μ~nω((Λ)𝒦L)eLn\tilde{\mu}_{n}^{\omega}(\mathcal{M}(\Lambda)\setminus\mathcal{K}_{L})\geq e^{-Ln} is finite for mpm_{p}-almost every ωΩ.\omega\in\Omega.

3.5. Proof of Theorem 1.5

We fix a metric on (Λ)\mathcal{M}(\Lambda) that generates the weak* topology, and a countable dense subset 𝒟\mathcal{D} of on (Λ)\mathcal{M}(\Lambda). For μ𝒟\mu\in\mathcal{D}, LL\in\mathbb{N} let B(μ,1/L)B(\mu,1/L) denote the closed ball of radius 1/L1/L about μ\mu. By Lemma 3.8, there exists a Borel subset Γ(B(μ,1/L))\Gamma(B(\mu,1/L)) of Ω\Omega with full mpm_{p}-measure such that if ωΓ(B(μ,1/L))\omega\in\Gamma(B(\mu,1/L)) then

(3.8) lim supn1nlogμ~nω(B(μ,1/L))infB(μ,1/L)Ip.\limsup_{n\to\infty}\frac{1}{n}\log\tilde{\mu}_{n}^{\omega}(B(\mu,1/L))\leq-\inf_{B(\mu,1/L)}I_{p}.

In view of Lemma 3.9, we fix an increasing sequence (𝒦L)L=1(\mathcal{K}_{L})_{L=1}^{\infty} of compact subsets of (Λ)\mathcal{M}(\Lambda) and a sequence (ΓL)L=1(\Gamma_{L})_{L=1}^{\infty} of Borel subsets of Ω\Omega with full mpm_{p}-measure such that L=1𝒦L=(Λ)\bigcup_{L=1}^{\infty}\mathcal{K}_{L}=\mathcal{M}(\Lambda), and for all LL\in\mathbb{N} and all ωΓL\omega\in\Gamma_{L},

(3.9) lim supn1nlogμ~nω((Λ)𝒦L)L.\limsup_{n\to\infty}\frac{1}{n}\log\tilde{\mu}^{\omega}_{n}(\mathcal{M}(\Lambda)\setminus\mathcal{K}_{L})\leq-L.

We set

Γ=(μ𝒟L=1Γ(B(μ,1/L)))(L=1ΓL).\Gamma=\left(\bigcap_{\mu\in\mathcal{D}}\bigcap_{L=1}^{\infty}\Gamma(B(\mu,1/L))\right)\cap\left(\bigcap_{L=1}^{\infty}\Gamma_{L}\right).

Clearly we have mp(Γ)=1m_{p}(\Gamma)=1. If ωΓ\omega\in\Gamma, then (μ~nω)n=1(\tilde{\mu}_{n}^{\omega})_{n=1}^{\infty} is exponentially tight by (3.9).

Let 𝒞\mathcal{C} be a non-empty closed subset of (Λ)\mathcal{M}(\Lambda) and let LL\in\mathbb{N}. Let 𝒢\mathcal{G} be an open subset of (Λ)\mathcal{M}(\Lambda) that contains 𝒞𝒦L\mathcal{C}\cap\mathcal{K}_{L}. Since 𝒞𝒦L\mathcal{C}\cap\mathcal{K}_{L} is compact, there exists a finite subset {μ1,,μs}\{\mu_{1},\ldots,\mu_{s}\} of 𝒟\mathcal{D} and L1,,LsL_{1},\ldots,L_{s}\in\mathbb{N} such that 𝒞𝒦Li=1sB(μi,1/Li)𝒢\mathcal{C}\cap\mathcal{K}_{L}\subset\bigcup_{i=1}^{s}B(\mu_{i},1/L_{i})\subset\mathcal{G}. By (3.8) applied to each of these closed balls, we have

lim supn1nlogμ~nω(𝒞𝒦L)max1islim supn1nlogμ~nω(B(μi,1/Li))max1is(infB(μi,1/Li)Ip)inf𝒢Ip.\begin{split}\limsup_{n\to\infty}\frac{1}{n}\log\tilde{\mu}_{n}^{\omega}(\mathcal{C}\cap\mathcal{K}_{L})&\leq\max_{1\leq i\leq s}\limsup_{n\to\infty}\frac{1}{n}\log\tilde{\mu}_{n}^{\omega}(B(\mu_{i},1/L_{i}))\\ &\leq\max_{1\leq i\leq s}\left(-\inf_{B(\mu_{i},1/L_{i})}I_{p}\right)\leq-\inf_{\mathcal{G}}I_{p}.\end{split}

Since 𝒢\mathcal{G} is an arbitrary open set containing 𝒞𝒦L\mathcal{C}\cap\mathcal{K}_{L} and IpI_{p} is lower semicontinuous,

(3.10) lim supn1nlogμ~nω(𝒞𝒦L)inf𝒞𝒦LIp.\limsup_{n\to\infty}\frac{1}{n}\log\tilde{\mu}_{n}^{\omega}(\mathcal{C}\cap\mathcal{K}_{L})\leq-\inf_{\mathcal{C}\cap\mathcal{K}_{L}}I_{p}.

From (3.9) and (3.10), for every ωΓ\omega\in\Gamma we obtain

(3.11) lim supn1nlogμ~nω(𝒞)max{inf𝒞𝒦LIp,L}.\limsup_{n\to\infty}\frac{1}{n}\log\tilde{\mu}^{\omega}_{n}(\mathcal{C})\leq\max\left\{-\inf_{\mathcal{C}\cap\mathcal{K}_{L}}I_{p},-L\right\}.

If Linf𝒞𝒦LIpL\geq\inf_{\mathcal{C}\cap\mathcal{K}_{L}}I_{p}, then (3.11) yields

lim supn1nlogμ~nω(𝒞)inf𝒞𝒦LIpinf𝒞Ip.\limsup_{n\to\infty}\frac{1}{n}\log\tilde{\mu}^{\omega}_{n}(\mathcal{C})\leq-\inf_{\mathcal{C}\cap\mathcal{K}_{L}}I_{p}\leq-\inf_{\mathcal{C}}I_{p}.

Combining this with (3.9) we obtain the desired inequality. If L<inf𝒞𝒦LIpL<\inf_{\mathcal{C}\cap\mathcal{K}_{L}}I_{p} for all LL\in\mathbb{N}, then we obtain inf𝒞Ip=\inf_{\mathcal{C}}I_{p}=\infty since (𝒦L)L=1(\mathcal{K}_{L})_{L=1}^{\infty} is increasing and L=1𝒦L=(Λ)\bigcup_{L=1}^{\infty}\mathcal{K}_{L}=\mathcal{M}(\Lambda). Moreover, (3.11) yields lim supn(1/n)logμ~nω(𝒞)=.\limsup_{n}(1/n)\log\tilde{\mu}^{\omega}_{n}(\mathcal{C})=-\infty. The proof of Theorem 1.5(a) is complete.

By Theorem 1.5(a), (μ~nω)n=1(\tilde{\mu}_{n}^{\omega})_{n=1}^{\infty} is tight for mpm_{p}-almost every ωΩ\omega\in\Omega. By Prohorov’s theorem, it has a limit point. Let (μ~njω)j=1(\tilde{\mu}_{n_{j}}^{\omega})_{j=1}^{\infty} be an arbitrary convergent subsequence of (μ~nω)n=1(\tilde{\mu}_{n}^{\omega})_{n=1}^{\infty} with the limit measure μ~ω\tilde{\mu}^{\omega}. For a proof of Theorem 1.5(b) it suffices to show μ~ω=δmpλp\tilde{\mu}^{\omega}=\delta_{m_{p}\otimes\lambda_{p}}.

We fix a metric that generates the weak* topology on (Λ)\mathcal{M}(\Lambda). Since IpI_{p} is a good rate function by Theorem 1.4(a), for any c>0c>0 the level set Ipc={μ(Λ):Ip(μ)c}I_{p}^{c}=\{\mu\in\mathcal{M}(\Lambda)\colon I_{p}(\mu)\leq c\} is compact. Let ν(Λ){mpλp}\nu\in\mathcal{M}(\Lambda)\setminus\{m_{p}\otimes\lambda_{p}\}. By the last assertion of Proposition 2.3 we have Ip(ν)>0I_{p}(\nu)>0, and so νIpI(ν)/2\nu\notin I_{p}^{I(\nu)/2}. Take r>0r>0 such that the closed ball B(ν,r)B(\nu,r) of radius rr about ν\nu in (Λ)\mathcal{M}(\Lambda) does not intersect IpI(ν)/2I_{p}^{I(\nu)/2}. By the weak* convergence of (μ~njω)j=1(\tilde{\mu}_{n_{j}}^{\omega})_{j=1}^{\infty} to μ~ω\tilde{\mu}^{\omega} and the large deviations upper bound for closed sets in Theorem 1.5(a), we have

μ~ω(int(B(ν,r)))lim infjμ~njω(int(B(ν,r)))lim supjμ~njω(B(ν,r))lim supjexp(Ip(ν)nj/2)=0.\begin{split}\tilde{\mu}^{\omega}({\rm int}(B(\nu,r)))&\leq\liminf_{j\to\infty}\tilde{\mu}_{n_{j}}^{\omega}({\rm int}(B(\nu,r)))\leq\limsup_{j\to\infty}\tilde{\mu}_{n_{j}}^{\omega}(B(\nu,r))\\ &\leq\limsup_{j\to\infty}\exp(-I_{p}(\nu)n_{j}/2)=0.\end{split}

Hence, the support of μ~ω\tilde{\mu}^{\omega} does not contain ν\nu. Since ν\nu is an arbitrary element of (Λ)\mathcal{M}(\Lambda) which is not mpλpm_{p}\otimes\lambda_{p}, it follows that μ~ω=δmpλp\tilde{\mu}^{\omega}=\delta_{m_{p}\otimes\lambda_{p}}. The proof of Theorem 1.5(b) is complete. ∎

Remark 3.10.

Since (Λ)\mathcal{M}(\Lambda) is non-compact, the tightness in Theorem 1.5(a) was used in establishing the convergence in Theorem 1.5(b). Nevertheless, (Ω×[0,1])\mathcal{M}(\Omega\times[0,1]) is compact. By applying the Contraction Principle to the inclusion (Λ)(Ω×[0,1])\mathcal{M}(\Lambda)\hookrightarrow\mathcal{M}(\Omega\times[0,1]), one can transfer the LDP in Theorem 1.4(a) to the LDP for the sequence (μ~n)n=1(\tilde{\mu}_{n})_{n=1}^{\infty} viewed as a sequence in 2(Ω×[0,1])\mathcal{M}^{2}(\Omega\times[0,1]). Using the latter LDP, one can establish a version of the upper bound in Theorem 1.5(a) for any closed subset of (Ω×[0,1])\mathcal{M}(\Omega\times[0,1]), as well as the convergence of (μ~n)n=1(\tilde{\mu}_{n})_{n=1}^{\infty} to δmpλp\delta_{m_{p}\otimes\lambda_{p}} in 2(Ω×[0,1])\mathcal{M}^{2}(\Omega\times[0,1]). These are actually sufficient for the proof of Theorem 1.1.

One merit of considering large deviations on the non-compact space (Λ)\mathcal{M}(\Lambda) rather than on (Ω×[0,1])\mathcal{M}(\Omega\times[0,1]) is that one can permit bounded continuous functions on Λ\Lambda that are naturally associated with the random continued fraction expansion (1.1), and do not have continuous extensions to Ω×[0,1]\Omega\times[0,1]. See Corollary 4.19 for details.

4. Establishing the LDP for the Gauss-Rényi map

This last section is mostly dedicated to the proof of Theorem 1.4. In §4.1 we summarize results on the thermodynamic formalism for the countable full shift. In §4.2 we consider an inducing scheme of the full shift and introduce a symbolic coding of the associated induced system. In §4.3 we recall the result of the second-named author [42] that give a sufficient condition for the level-2 LDP on periodic points in terms of induced potentials. We also recall the result in [40] on the uniqueness of minimizer of the rate function. In order to implement all these results, in §\S4.4 we show that the Gauss-Rényi map is topologically conjugate to the shift map on the countable full shift. In §4.5 we perform distortion estimates for an induced version of the annealed geometric potential φ\varphi. In §4.6 we establish the existence and uniqueness of the equilibrium state for the symbolic version of the potential φ\varphi, and show that this equilibrium state is the symbolic version of the measure mpλpm_{p}\otimes\lambda_{p}. In §4.7 we complete the proof of Theorem 1.4. In §4.8 we state two corollaries of independent interest on annealed and quenched level-1 large deviations, and apply them to the problem of frequency of digits in the random continued fraction expansion.

4.1. Thermodynamic formalism for the countable full shift

Consider the countable full shift

(4.1) ={z=(zn)n=1:zn for n},\mathbb{N}^{\mathbb{N}}=\{z=(z_{n})_{n=1}^{\infty}\colon z_{n}\in\mathbb{N}\text{ for }n\in\mathbb{N}\},

which is the cartesian product topological space of the discrete space \mathbb{N}. We introduce main constituent components of the thermodynamic formalism for the countable full shift (4.1), and state a variational principle and a relationship between equilibrium states and Gibbs states. Our main reference is [27] that contains results on countable Markov shifts which are not necessarily the full shift.

The left shift σ:\sigma\colon\mathbb{N}^{\mathbb{N}}\to\mathbb{N}^{\mathbb{N}} given by σ(zn)n=1=(zn+1)n=1\sigma(z_{n})_{n=1}^{\infty}=(z_{n+1})_{n=1}^{\infty} is continuous. For nn\in\mathbb{N} and a1anna_{1}\cdots a_{n}\in\mathbb{N}^{n}, define an nn-cylinder

[a1an]={z:zi=ai for i=1,,n}.[a_{1}\cdots a_{n}]=\{z\in\mathbb{N}^{\mathbb{N}}\colon z_{i}=a_{i}\text{ for }i=1,\ldots,n\}.

Let (,σ)\mathcal{M}(\mathbb{N}^{\mathbb{N}},\sigma) denote the set of σ\sigma-invariant Borel probability measures. For each μ(,σ)\mu\in\mathcal{M}(\mathbb{N}^{\mathbb{N}},\sigma), let h(μ)[0,]h(\mu)\in[0,\infty] denote the measure-theoretic entropy of μ\mu with respect to σ\sigma. Let ϕ:\phi\colon\mathbb{N}^{\mathbb{N}}\to\mathbb{R} be a function, called a potential. For each nn\in\mathbb{N} we write SnϕS_{n}\phi for the Birkhoff sum i=0n1ϕσi\sum_{i=0}^{n-1}\phi\circ\sigma^{i}, and introduce a pressure

P(ϕ)=limn1nloga1annsup[a1an]expSnϕ.P(\phi)=\lim_{n\to\infty}\frac{1}{n}\log\sum_{a_{1}\cdots a_{n}\in\mathbb{N}^{n}}\sup_{[a_{1}\cdots a_{n}]}\exp S_{n}\phi.

This limit exists by the sub-additivity, which is never -\infty. We say:

  • ϕ\phi is acceptable if it is uniformly continuous and satisfies

    supa(sup[a]ϕinf[a]ϕ)<;\sup_{a\in\mathbb{N}}\left(\sup_{[a]}\phi-\inf_{[a]}\phi\right)<\infty;
  • ϕ\phi is locally Hölder continuous if there exist constants K>0K>0 and γ(0,1)\gamma\in(0,1) such that varn(ϕ)Kγn{\rm var}_{n}(\phi)\leq K\gamma^{n}, where

    varn(ϕ)=sup{ϕ(z)ϕ(w):z,w,zi=wi for i=1,,n}.{\rm var}_{n}(\phi)=\sup\{\phi(z)-\phi(w)\colon z,w\in\mathbb{N}^{\mathbb{N}},\ z_{i}=w_{i}\ \text{ for }i=1,\ldots,n\}.

Let ϕ:\phi\colon\mathbb{N}^{\mathbb{N}}\to\mathbb{R} be acceptable and satisfy P(ϕ)<P(\phi)<\infty. Then supϕ\sup\phi is finite (see [27, Proposition 2.1.9]). Let

ϕ(,σ)={μ(,σ):ϕ𝑑μ>}.\mathcal{M}_{\phi}(\mathbb{N}^{\mathbb{N}},\sigma)=\left\{\mu\in\mathcal{M}(\mathbb{N}^{\mathbb{N}},\sigma)\colon\int\phi d\mu>-\infty\right\}.

By [27, Theorem 2.1.7], for any μϕ(,σ)\mu\in\mathcal{M}_{\phi}(\mathbb{N}^{\mathbb{N}},\sigma) we have h(μ)+ϕ𝑑μP(ϕ)<h(\mu)+\int\phi d\mu\leq P(\phi)<\infty, and so h(μ)<h(\mu)<\infty. The following equality is known as the variational principle.

Proposition 4.1 ([27, Theorem 2.1.7, Theorem 2.1.8]).

Let ϕ:\phi\colon\mathbb{N}^{\mathbb{N}}\to\mathbb{R} be acceptable and satisfy P(ϕ)<P(\phi)<\infty. Then

P(ϕ)=sup{h(μ)+ϕ𝑑μ:μϕ(,σ)}.P(\phi)=\sup\left\{h(\mu)+\int\phi d\mu\colon\mu\in\mathcal{M}_{\phi}(\mathbb{N}^{\mathbb{N}},\sigma)\right\}.

Let ϕ:\phi\colon\mathbb{N}^{\mathbb{N}}\to\mathbb{R} be acceptable and satisfy P(ϕ)<P(\phi)<\infty. A measure μϕ(,σ)\mu\in\mathcal{M}_{\phi}(\mathbb{N}^{\mathbb{N}},\sigma) is called an equilibrium state for the potential ϕ\phi if

P(ϕ)=h(μ)+ϕ𝑑μ.P(\phi)=h(\mu)+\int\phi d\mu.

A measure μ()\mu\in\mathcal{M}(\mathbb{N}^{\mathbb{N}}) is called a Gibbs state for the potential ϕ\phi if there exists a constant K1K\geq 1 such that for all nn\in\mathbb{N}, all a1anna_{1}\cdots a_{n}\in\mathbb{N}^{n} and all x[a1an]x\in[a_{1}\cdots a_{n}],

K1μ([a1an])exp(Snϕ(x)P(ϕ)n)K.K^{-1}\leq\frac{\mu([a_{1}\cdots a_{n}])}{\exp(S_{n}\phi(x)-P(\phi)n)}\leq K.
Proposition 4.2 ([27, Theorem 2.2.9, Corollary 2.7.5]).

Let ϕ:\phi\colon\mathbb{N}^{\mathbb{N}}\to\mathbb{R} be locally Hölder continuous and satisfy P(ϕ)<P(\phi)<\infty. Then there exists a unique shift-invariant Gibbs state μϕ\mu_{\phi} for ϕ\phi. If ϕ𝑑μϕ>\int\phi d\mu_{\phi}>-\infty, then μϕ\mu_{\phi} is the unique equilibrium state for ϕ\phi.

4.2. Coding of the induced system

Consider the inducing scheme ([1],t[1])(\mathbb{N}^{\mathbb{N}}\setminus[1],t_{\mathbb{N}^{\mathbb{N}}\setminus[1]}) of the left shift σ:\sigma\colon\mathbb{N}^{\mathbb{N}}\to\mathbb{N}^{\mathbb{N}}. We show that the associated induced system σ^:^^\widehat{\sigma}\colon{\widehat{\mathbb{N}}}^{\mathbb{N}}\to{\widehat{\mathbb{N}}}^{\mathbb{N}} is in a natural way topologically conjugate to the full shift over an infinite alphabet.

We introduce the empty word \emptyset by the rule ω=ω=ω\omega\emptyset=\omega=\emptyset\omega for any word ω\omega from \mathbb{N}. For each nn\in\mathbb{N}, write 1n1^{n} for 111n11\cdots 1\in\mathbb{N}^{n}, the nn-string of 11. We set 10=1^{0}=\emptyset for convenience. We introduce an infinite alphabet

(4.2) 𝕄={b{1}[a1nb]:a{1} and n{0}},\mathbb{M}=\left\{\bigcup_{b\in\mathbb{N}\setminus\{1\}}[a1^{n}b]\colon a\in\mathbb{N}\setminus\{1\}\text{ and }n\in\mathbb{N}\cup\{0\}\right\},

which is a collection of pairwise disjoint subsets of [1]\mathbb{N}^{\mathbb{N}}\setminus[1]. We endow 𝕄\mathbb{M} with the discrete topology, and introduce the countable full shift

(4.3) 𝕄={(xn)n=1:xn𝕄 for n},{\mathbb{M}}^{\mathbb{N}}=\{(x_{n})_{n=1}^{\infty}\colon x_{n}\in\mathbb{M}\text{ for }n\in\mathbb{N}\},

which is the cartesian product topological space of 𝕄\mathbb{M}. Clearly 𝕄\mathbb{M}^{\mathbb{N}} is topologically isomorphic to \mathbb{N}^{\mathbb{N}}. With a slight abuse of notation let σ:𝕄𝕄\sigma\colon{\mathbb{M}}^{\mathbb{N}}\to{\mathbb{M}}^{\mathbb{N}} denote the left shift.

We define a map ι:𝕄^\iota\colon\mathbb{M}^{\mathbb{N}}\to{\widehat{\mathbb{N}}}^{\mathbb{N}} as follows. Let (xn)n=1𝕄(x_{n})_{n=1}^{\infty}\in\mathbb{M}^{\mathbb{N}}. By the definition of 𝕄\mathbb{M} in (4.2), for every nn\in\mathbb{N} we have xn=b{1}[an1jnb]x_{n}=\bigcup_{b\in\mathbb{N}\setminus\{1\}}[a_{n}1^{j_{n}}b] where an{1}a_{n}\in\mathbb{N}\setminus\{1\} and jn{0}j_{n}\in\mathbb{N}\cup\{0\}. We set

ι((xn)n=1)n=1[a11j1a21j2an1jn].\iota((x_{n})_{n=1}^{\infty})\in\bigcap_{n=1}^{\infty}[a_{1}1^{j_{1}}a_{2}1^{j_{2}}\cdots a_{n}1^{j_{n}}].
Lemma 4.3.

The map ι\iota is a homeomorphism, and satisfies ισ=σ^ι\iota\circ\sigma=\widehat{\sigma}\circ\iota.

Proof.

Clearly ι\iota is continuous and injective. For every a{1}a\in\mathbb{N}\setminus\{1\} and every n{0}n\in\mathbb{N}\cup\{0\}, the set b{1}[a1nb]\bigcup_{b\in\mathbb{N}\setminus\{1\}}[a1^{n}b] is mapped by σ^\widehat{\sigma} bijectively onto [1]\mathbb{N}^{\mathbb{N}}\setminus[1]. Moreover, the collection of sets of this form defines a partition of the set k=1{t=k}\bigcup_{k=1}^{\infty}\{t=k\}, namely

k=1{t=k}=a{1}n{0}b{1}[a1nb].\bigcup_{k=1}^{\infty}\{t=k\}=\bigcup_{a\in\mathbb{N}\setminus\{1\}}\bigcup_{n\in\mathbb{N}\cup\{0\}}\bigcup_{b\in\mathbb{N}\setminus\{1\}}[a1^{n}b].

All the unions are disjoint unions. It follows that ι(𝕄)=^\iota(\mathbb{M}^{\mathbb{N}})=\widehat{\mathbb{N}}^{\mathbb{N}}. The last assertion follows from the definition of ι.\iota.

4.3. Level-2 LDP for the countable full shift

Let ϕ:\phi\colon\mathbb{N}^{\mathbb{N}}\to\mathbb{R} be acceptable and satisfy P(ϕ)<P(\phi)<\infty. We are concerned with the LDP a sequence (ν~n)n=1(\tilde{\nu}_{n})_{n=1}^{\infty} of Borel probability measures on ()\mathcal{M}(\mathbb{N}^{\mathbb{N}}) given by

(4.4) ν~n=1Zn(ϕ)xFix(σn)exp(Snϕ(x))δVnσ(x),\tilde{\nu}_{n}=\frac{1}{Z_{n}(\phi)}\sum_{x\in{\rm Fix}(\sigma^{n})}\exp(S_{n}\phi(x))\delta_{V_{n}^{\sigma}(x)},

where Vnσ(x)()V_{n}^{\sigma}(x)\in\mathcal{M}(\mathbb{N}^{\mathbb{N}}) denotes the uniform probability distribution on the orbit (σix)i=0n1(\sigma^{i}x)_{i=0}^{n-1}, and δVnσ(x)\delta_{V_{n}^{\sigma}(x)} denotes the Borel probability measure on ()\mathcal{M}(\mathbb{N}^{\mathbb{N}}) that is the unit point mass at Vnσ(x)V_{n}^{\sigma}(x), and Zn(ϕ)Z_{n}(\phi) denotes the normalizing constant. We introduce a free energy Fϕ:()[,0]F_{\phi}\colon\mathcal{M}(\mathbb{N}^{\mathbb{N}})\to[-\infty,0] by

Fϕ(μ)={h(μ)+ϕ𝑑μ if μϕ(,σ), otherwise.F_{\phi}(\mu)=\begin{cases}h(\mu)+\int\phi d\mu&\text{ if $\mu\in\mathcal{M}_{\phi}(\mathbb{N}^{\mathbb{N}},\sigma)$},\\ -\infty&\text{ otherwise.}\end{cases}

The function Fϕ+P(ϕ)-F_{\phi}+P(\phi) is a natural candidate for the rate function of this LDP. However, this function may not be lower semicontinuous since the entropy function is not upper semicontinuous. Hence, we take the lower semicontinuous regularization of Fϕ+P(ϕ)-F_{\phi}+P(\phi). Define Iϕ:()[0,]I_{\phi}\colon\mathcal{M}(\mathbb{N}^{\mathbb{N}})\to[0,\infty] by

(4.5) Iϕ(μ)=inf𝒢μsupν𝒢Fϕ(ν)+P(ϕ),I_{\phi}(\mu)=-\inf_{\mathcal{G}\ni\mu}\sup_{\nu\in\mathcal{G}}F_{\phi}(\nu)+P(\phi),

where the supremum is taken over all measures in an open subset 𝒢\mathcal{G} of ()\mathcal{M}(\mathbb{N}^{\mathbb{N}}) that contains μ\mu, and the infimum is taken over all such open subsets. Then IϕI_{\phi} is lower semicontinuous and satisfies IϕFϕ+P(ϕ)I_{\phi}\leq-F_{\phi}+P(\phi).

If there is a Gibbs state for the potential ϕ\phi, then the LDP holds for (ν~n)n=1(\tilde{\nu}_{n})_{n=1}^{\infty} from the result in [38]. Due to the existence of the neutral fixed point of the Rényi map T1T_{1}, the annealed Gauss-Rényi measure ηp\eta_{p} is not a Gibbs state for the potential ψ\psi (see Lemma 4.12). Hence [38] cannot be applied to (,ψ)(\mathbb{N}^{\mathbb{N}},\psi). Instead we apply the result in [42] on the LDP for (ν~n)n=1(\tilde{\nu}_{n})_{n=1}^{\infty} when a Gibbs state for ϕ\phi does not exist.

Using the conjugacy ι\iota in §4.2, we introduce a parametrized family of twisted induced potentials Φγ:𝕄\Phi_{\gamma}\colon{\mathbb{M}}^{\mathbb{N}}\to\mathbb{R} (γ\gamma\in\mathbb{R}) by

(4.6) Φγ(ι(x))=St[1](ι(x))ϕ(ι(x))γt[1](ι(x)).\Phi_{\gamma}(\iota(x))=S_{t_{\mathbb{N}^{\mathbb{N}}\setminus[1]}(\iota(x))}\phi(\iota(x))-\gamma t_{\mathbb{N}^{\mathbb{N}}\setminus[1]}(\iota(x)).
Theorem 4.4 ([42, Theorem A]).

Let ϕ:\phi\colon\mathbb{N}^{\mathbb{N}}\to\mathbb{R} be acceptable and satisfy P(ϕ)<P(\phi)<\infty. Suppose the twisted induced potentials Φγ:𝕄\Phi_{\gamma}\colon\mathbb{M}^{\mathbb{N}}\to\mathbb{R} (γ)(\gamma\in\mathbb{R}) are locally Hölder continuous, and there exists γ0\gamma_{0}\in\mathbb{R} such that P(Φγ0)=0P(\Phi_{\gamma_{0}})=0. Then (ν~n)n=1(\tilde{\nu}_{n})_{n=1}^{\infty} is exponentially tight and satisfies the LDP with the good rate function IϕI_{\phi}.

The uniqueness of minimizer of the rate function IϕI_{\phi} does not follow from Theorem 4.4 and should be examined on a case-by-case basis. An ideal situation is that the shift-invariant Gibbs state for ϕ\phi is unique, the equilibrium state for ϕ\phi is unique, the minimizer of IϕI_{\phi} is unique, and all these three coincide. However this is not always the case. Under the hypothesis of Theorem 4.4, by virtue of Proposition 4.2 there exists a unique Gibbs state for the potential ϕ\phi. If moreover ϕ\phi is integrable against the Gibbs state, then it is the unique equilibrium state for ϕ\phi, and clearly is a minimizer of IϕI_{\phi}. Conversely, a minimizer of IϕI_{\phi} may not be an equilibrium state for ϕ\phi in general: an example of a potential ϕ:\phi\colon\mathbb{N}^{\mathbb{N}}\to\mathbb{R} can be found in [35] for which there is a Gibbs state μ(,σ)\mu\in\mathcal{M}(\mathbb{N}^{\mathbb{N}},\sigma) such that Iϕ(μ)=0I_{\phi}(\mu)=0 and μ\mu is not an equilibrium state since ϕ𝑑μ=\int\phi d\mu=-\infty.

Under additional hypothesis on the potential, one can show that any minimizer is an equilibrium state. We say ϕ:\phi\colon\mathbb{N}^{\mathbb{N}}\to\mathbb{R} is summable if ksup[k]eϕ\sum_{k\in\mathbb{N}}\sup_{[k]}e^{\phi} is finite. If ϕ\phi is summable, then P(ϕ)<P(\phi)<\infty. Set

β(ϕ)=inf{β:βϕ is summable}.\beta_{\infty}(\phi)=\inf\left\{\beta\in\mathbb{R}\colon\text{$\beta\phi$ is summable}\right\}.
Proposition 4.5.

Let ϕ:\phi\colon\mathbb{N}^{\mathbb{N}}\to\mathbb{R} be uniformly continuous and summable with β(ϕ)<1\beta_{\infty}(\phi)<1. Then, any minimizer of IϕI_{\phi} is an equilibrium state for the potential ϕ\phi.

A proof of this proposition is briefly outline as follows. By the definition (4.5), if μ\mu is a minimizer of IϕI_{\phi} then there is a sequence (μk)k=1(\mu_{k})_{k=1}^{\infty} in ϕ(,σ)\mathcal{M}_{\phi}(\mathbb{N}^{\mathbb{N}},\sigma) that converges to μ\mu in the weak* topology with limkFϕ(μk)=0\lim_{k}F_{\phi}(\mu_{k})=0. Based on this information we show that μ\mu is an equilibrium state for ϕ\phi. The case limkh(μk)=0\lim_{k}h(\mu_{k})=0 is easy to handle, while the case limkh(μk)=\lim_{k}h(\mu_{k})=\infty (and hence limkϕ𝑑μk\lim_{k}\int\phi d\mu_{k}\to-\infty) requires attention. A key ingredient in the latter case is the upper semicontinuity of the map μkh(μk)/(ϕ𝑑μk)\mu_{k}\mapsto h(\mu_{k})/(-\int\phi d\mu_{k}), as proved in [40, Theorem 2.4] inspired by [14, Lemma 6.5].

Proof of Proposition 4.5.

The following proof is almost a repetition of the proof of [40, Theorem 2.1] for the reader’s convenience. Considering ϕP(ϕ)\phi-P(\phi) instead of ϕ\phi, we may assume P(ϕ)=0P(\phi)=0. Let μ(,σ)\mu\in\mathcal{M}(\mathbb{N}^{\mathbb{N}},\sigma) be a minimizer of IϕI_{\phi}. Since (,σ)\mathcal{M}(\mathbb{N}^{\mathbb{N}},\sigma) is a closed subset of (,σ)\mathcal{M}(\mathbb{N}^{\mathbb{N}},\sigma), μ\mu is shift-invariant. By the definition (4.5), there is a sequence (μk)k=1(\mu_{k})_{k=1}^{\infty} in ϕ(,σ)\mathcal{M}_{\phi}(\mathbb{N}^{\mathbb{N}},\sigma) that converges to μ\mu in the weak* topology with limkFϕ(μk)=0\lim_{k}F_{\phi}(\mu_{k})=0. By [40, Lemma 2.3], we have infkϕ𝑑μk>\inf_{k}\int\phi d\mu_{k}>-\infty. By this and supϕ<\sup\phi<\infty, a simple upper semicontinuity argument as in [40, Remark 2.5] shows ϕ𝑑μ>\int\phi d\mu>-\infty. If lim infkh(μk)=0\liminf_{k}h(\mu_{k})=0, then for any subsequence (μkj)j=1(\mu_{k_{j}})_{j=1}^{\infty} with limjh(μkj)=0\lim_{j}h(\mu_{k_{j}})=0 we have

0=limjFϕ(μkj)ϕ𝑑μh(μ)+ϕ𝑑μ=Fϕ(μ).0=\lim_{j\to\infty}F_{\phi}(\mu_{k_{j}})\leq\int\phi d\mu\leq h(\mu)+\int\phi d\mu=F_{\phi}(\mu).

Since Fϕ(μ)P(ϕ)=0F_{\phi}(\mu)\leq P(\phi)=0, μ\mu is an equilibrium state for ϕ\phi. If lim infkh(μk)>0\liminf_{k}h(\mu_{k})>0, then we have lim infk(ϕ𝑑μk)>0\liminf_{k}(-\int\phi d\mu_{k})>0 and

0=limkFϕ(μk)=limk(ϕ𝑑μk)(h(μk)ϕ𝑑μk1).0=\lim_{k\to\infty}F_{\phi}(\mu_{k})=\lim_{k\to\infty}\left(-\int\phi d\mu_{k}\right)\left(\frac{h(\mu_{k})}{-\int\phi d\mu_{k}}-1\right).

It follows that

limk(h(μk)ϕ𝑑μk1)=0.\lim_{k\to\infty}\left(\frac{h(\mu_{k})}{-\int\phi d\mu_{k}}-1\right)=0.

We have ϕ𝑑μh(μ)-\int\phi d\mu\geq h(\mu). If ϕ𝑑μ=0-\int\phi d\mu=0, then clearly μ\mu is an equilibrium state for ϕ\phi. If ϕ𝑑μ>0-\int\phi d\mu>0, then by [40, Theorem 2.4] we have

h(μ)ϕ𝑑μ10,\frac{h(\mu)}{-\int\phi d\mu}-1\geq 0,

namely Fϕ(μ)0F_{\phi}(\mu)\geq 0. Since Fϕ(μ)0F_{\phi}(\mu)\leq 0, μ\mu is an equilibrium state for ϕ\phi. The proof of Proposition 4.5 is complete. ∎

4.4. Symbolic coding of the Gauss-Rényi map

The next proposition allows us to introduce a symbolic representation of the Gauss-Rényi map.

Proposition 4.6.

The following statements hold.

  • (a)

    For every (an)n(a_{n})_{n\in\mathbb{N}}\in\mathbb{N}^{\mathbb{N}} we have n=1Δ(a1an)={(ω,x)}Λ\bigcap_{n=1}^{\infty}\varDelta(a_{1}\cdots a_{n})=\{(\omega,x)\}\subset\Lambda, where ωnanmod2\omega_{n}\equiv a_{n}\mod 2, Cn=(an+ωn)/2+ωn+1C_{n}=(a_{n}+\omega_{n})/2+\omega_{n+1} and

    x=ω1+(1)ω1  C1+(1)ω2  C2+(1)ω3  C3+.x=\omega_{1}+\frac{\displaystyle{\hfill{(-1)^{\omega_{1}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{1}}\hfill}}+\frac{\displaystyle{\hfill{(-1)^{\omega_{2}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{2}}\hfill}}+\frac{\displaystyle{\hfill{(-1)^{\omega_{3}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{3}}\hfill}}+\cdots.
  • (b)

    For every (ω,x)Λ(\omega,x)\in\Lambda we have {(ω,x)}=n=1Δ(a1an)\{(\omega,x)\}=\bigcap_{n=1}^{\infty}\varDelta(a_{1}\cdots a_{n}), where an=2Cn(ω,x)+ωn2ωn+1a_{n}=2C_{n}(\omega,x)+\omega_{n}-2\omega_{n+1}.

Proof.

As for (a), let (an)n(a_{n})_{n\in\mathbb{N}}\in\mathbb{N}^{\mathbb{N}}. Define (ωn)n{0,1}(\omega_{n})_{n\in\mathbb{N}}\in\{0,1\}^{\mathbb{N}} by ωnanmod2\omega_{n}\equiv a_{n}\mod 2, and Cn=(an+ωn)/2+ωn+1C_{n}=(a_{n}+\omega_{n})/2+\omega_{n+1} for nn\in\mathbb{N}. Note that (1)ωn+1+Cn1(-1)^{\omega_{n+1}}+C_{n}\geq 1 for every nn\in\mathbb{N}. By Lemma 2.2, the displayed continued fraction converges to a number x[0,1]x\in[0,1], and thus (ω,x)n=1Δ(a1an)(\omega,x)\in\bigcap_{n=1}^{\infty}\varDelta(a_{1}\cdots a_{n}). The algorithm described in §2.1 shows {(ω,x)}=n=1Δ(a1an)\{(\omega,x)\}=\bigcap_{n=1}^{\infty}\varDelta(a_{1}\cdots a_{n}). Since Rn(ω,x)=(θnω,Tωnx)R^{n}(\omega,x)=(\theta^{n}\omega,T_{\omega}^{n}x) we have

Tωnx=ωn+1+(1)ωn+1  Cn+1+(1)ωn+2  Cn+2+(1)ωn+3  Cn+3+.T_{\omega}^{n}x=\omega_{n+1}+\frac{\displaystyle{\hfill{(-1)^{\omega_{n+1}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{n+1}}\hfill}}+\frac{\displaystyle{\hfill{(-1)^{\omega_{n+2}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{n+2}}\hfill}}+\frac{\displaystyle{\hfill{(-1)^{\omega_{n+3}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{n+3}}\hfill}}+\cdots.

Hence (ω,x)Λ(\omega,x)\in\Lambda holds.

To prove (b), let (ω,x)Λ(\omega,x)\in\Lambda. Define an=2Cn(ω,x)ωn2ωn+1a_{n}=2C_{n}(\omega,x)-\omega_{n}-2\omega_{n+1} for nn\in\mathbb{N}. We have (1)ωn+1+Cn(ω,x)1(-1)^{\omega_{n+1}}+C_{n}(\omega,x)\geq 1 for every nn\in\mathbb{N}. Proposition 2.1(a) gives

x=ω1+(1)ω1  C1(ω,x)+(1)ω2  C2(ω,x)+(1)ω3  C3(ω,x)+,x=\omega_{1}+\frac{\displaystyle{\hfill{(-1)^{\omega_{1}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{1}(\omega,x)}\hfill}}+\frac{\displaystyle{\hfill{(-1)^{\omega_{2}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{2}(\omega,x)}\hfill}}+\frac{\displaystyle{\hfill{(-1)^{\omega_{3}}}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{C_{3}(\omega,x)}\hfill}}+\cdots,

which implies (ω,x)n=1Δ(a1an)(\omega,x)\in\bigcap_{n=1}^{\infty}\varDelta(a_{1}\cdots a_{n}). Proposition 4.6(a) yields {(ω,x)}=n=1Δ(a1an)\{(\omega,x)\}=\bigcap_{n=1}^{\infty}\varDelta(a_{1}\cdots a_{n}). ∎

Define a coding map π:Λ\pi\colon\mathbb{N}^{\mathbb{N}}\to\Lambda by

(4.7) π((zn)n=1)n=1Δ(z1zn).\pi((z_{n})_{n=1}^{\infty})\in\bigcap_{n=1}^{\infty}\varDelta(z_{1}\cdots z_{n}).

By Proposition 4.6, π\pi is well-defined and surjective. Obviously π\pi is continuous, injective and satisfies Rπ=πσR\circ\pi=\pi\circ\sigma. It is not hard to show that π\pi maps Borel sets to Borel sets. We set

(4.8) ηp=(mpλp)π,\eta_{p}=(m_{p}\otimes\lambda_{p})\circ\pi,

and call ηp\eta_{p} the annealed Gauss-Rényi measure. From (b) and (c) in Proposition 2.1, we have Λω=(0,1)\Lambda_{\omega}=(0,1)\setminus\mathbb{Q} for every ωΩ0\omega\in\Omega_{0}. This implies Ω0×((0,1))Λ\Omega_{0}\times((0,1)\setminus\mathbb{Q})\subset\Lambda, and so (mpλp)(Λ)=1(m_{p}\otimes\lambda_{p})(\Lambda)=1. Hence ηp\eta_{p} is a probability. The measure mpλpm_{p}\otimes\lambda_{p} is RR-invariant [23, Theorem 3.2] and by [23, Theorem 3.3] it is mixing. Hence ηp\eta_{p} is σ\sigma-invariant and mixing.

By Lemma 4.3, the induced system σ^:^^\widehat{\sigma}\colon{\widehat{\mathbb{N}}}^{\mathbb{N}}\to{\widehat{\mathbb{N}}}^{\mathbb{N}} is topologically conjugate to σ:𝕄𝕄\sigma\colon\mathbb{M}^{\mathbb{N}}\to\mathbb{M}^{\mathbb{N}} via ι\iota. Since R:ΛΛR\colon\Lambda\to\Lambda is topologically conjugate to σ:\sigma\colon\mathbb{N}^{\mathbb{N}}\to\mathbb{N}^{\mathbb{N}} via π\pi, the two induced systems R^:Λ^Λ^\widehat{R}\colon\widehat{\Lambda}\to\widehat{\Lambda} and σ^:^^\widehat{\sigma}\colon{\widehat{\mathbb{N}}}^{\mathbb{N}}\to{\widehat{\mathbb{N}}}^{\mathbb{N}} are topologically conjugate via π\pi. The three dynamical systems are summarized in the following diagram.

(4.9) 𝕄σ𝕄ιι^σ^^ππΛ^R^Λ^\begin{CD}\mathbb{M}^{\mathbb{N}}@>{\sigma}>{}>\mathbb{M}^{\mathbb{N}}\\ @V{\iota}V{}V@V{}V{\iota}V\\ {\widehat{\mathbb{N}}}^{\mathbb{N}}@>{\widehat{\sigma}}>{}>{\widehat{\mathbb{N}}}^{\mathbb{N}}\\ @V{\pi}V{}V@V{}V{\pi}V\\ \widehat{\Lambda}@>{\widehat{R}}>{}>\widehat{\Lambda}\\ \end{CD}

4.5. Refined distortion estimates

The distortion estimate in Lemma 3.4 does not suffice when a1ana_{1}\cdots a_{n} contains a long block of 11 that contains ana_{n}. The next lemma provides refined estimates in this case.

Lemma 4.7.

There exists a constant K>0K>0 such that if nn\in\mathbb{N}, ai=1a_{i}=1 for i=1,,ni=1,\ldots,n and an+11a_{n+1}\neq 1 then for any pair (ω,x),(ϱ,y)(\omega,x),(\varrho,y) of points in Δ(a1an+1)\varDelta(a_{1}\cdots a_{n+1}),

Snφ(ω,x)Snφ(ϱ,y){K|TωnxTϱny| if an+11,K|TωnxTϱny|12 if an+10.S_{n}\varphi(\omega,x)-S_{n}\varphi(\varrho,y)\leq\begin{cases}K|T_{\omega}^{n}x-T_{\varrho}^{n}y|&\text{ if }a_{n+1}\in\mathbb{N}_{1},\\ K|T_{\omega}^{n}x-T_{\varrho}^{n}y|^{\frac{1}{2}}&\text{ if }a_{n+1}\in\mathbb{N}_{0}.\end{cases}
Proof.

Let nn\in\mathbb{N} and suppose ai=1a_{i}=1 for i=1,,ni=1,\ldots,n and an+11a_{n+1}\neq 1. For i=0,,ni=0,\ldots,n put

qi={1i+2 if an+11,22i+an+1 if an+10,q_{i}=\begin{cases}\vskip 3.69885pt\displaystyle{\frac{1}{i+2}}&\text{ if }a_{n+1}\in\mathbb{N}_{1},\\ \displaystyle{\frac{2}{2i+a_{n+1}}}&\text{ if }a_{n+1}\in\mathbb{N}_{0},\end{cases}

and Ji=[qi+1,qi)J_{i}=\left[q_{i+1},q_{i}\right). Let (ω,x),(ϱ,y)Δ(a1an+1)(\omega,x),(\varrho,y)\in\varDelta(a_{1}\cdots a_{n+1}). We have T1(qi+1)=qiT_{1}(q_{i+1})=q_{i} for i=0,,n1i=0,\ldots,n-1 and x,yJn1x,y\in J_{n-1}. If an+11a_{n+1}\in\mathbb{N}_{1} then by Lemma 4.8 below applied to f=T1|[0,1/2)f=T_{1}|[0,1/2), there exists a uniform constant K1>0K_{1}>0 such that

(4.10) Snφ(ω,x)Snφ(ϱ,y)K1|TωnxTϱny|.S_{n}\varphi(\omega,x)-S_{n}\varphi(\varrho,y)\leq K_{1}|T_{\omega}^{n}x-T_{\varrho}^{n}y|.

If an+10a_{n+1}\in\mathbb{N}_{0} then we have

(4.11) |J0|=4an+12+2an+1 and i=0n1|Ji|2an+1.|J_{0}|=\frac{4}{a_{n+1}^{2}+2a_{n+1}}\ \text{ and }\ \sum_{i=0}^{n-1}|J_{i}|\leq\frac{2}{a_{n+1}}.

By Lemma 4.8 below applied to the restriction f=T1|[0,2/an+1)f=T_{1}|_{[0,2/a_{n+1})}, there exists a uniform constant K2>0K_{2}>0 such that

Snφ(ω,x)Snφ(ϱ,y)K2|TωnxTϱny||J0|i=0n1|Ji|.S_{n}\varphi(\omega,x)-S_{n}\varphi(\varrho,y)\leq K_{2}\frac{|T_{\omega}^{n}x-T_{\varrho}^{n}y|}{|J_{0}|}\sum_{i=0}^{n-1}|J_{i}|.

Since Rn(ω,x),Rn(ϱ,y)Δ(an+1)R^{n}(\omega,x),R^{n}(\varrho,y)\in\varDelta(a_{n+1}), the points TωnxT_{\omega}^{n}x, TϱnyT_{\varrho}^{n}y belong to the closure of J0J_{0}, and thus |TωnxTϱny|/|J0|1|T_{\omega}^{n}x-T_{\varrho}^{n}y|/|J_{0}|\leq 1. By this and (4.11),

(4.12) Snφ(ω,x)Snφ(ϱ,y)K2|TωnxTϱny||J0|i=0n1|Ji|K2|TωnxTϱny|12|J0|12i=0n1|Ji|K2an+12+2an+1an+1|TωnxTϱny|122K2|TωnxTϱny|12.\begin{split}S_{n}\varphi(\omega,x)-S_{n}\varphi(\varrho,y)&\leq K_{2}\frac{|T_{\omega}^{n}x-T_{\varrho}^{n}y|}{|J_{0}|}\sum_{i=0}^{n-1}|J_{i}|\\ &\leq K_{2}\frac{|T_{\omega}^{n}x-T_{\varrho}^{n}y|^{\frac{1}{2}}}{|J_{0}|^{\frac{1}{2}}}\sum_{i=0}^{n-1}|J_{i}|\\ &\leq K_{2}\frac{\sqrt{a_{n+1}^{2}+2a_{n+1}}}{a_{n+1}}|T_{\omega}^{n}x-T_{\varrho}^{n}y|^{\frac{1}{2}}\\ &\leq\sqrt{2}K_{2}|T_{\omega}^{n}x-T_{\varrho}^{n}y|^{\frac{1}{2}}.\end{split}

By (4.10) and (4.12), taking K=max{K1,2K2}K=\max\{K_{1},\sqrt{2}K_{2}\} yields the desired inequalities. ∎

The next general lemma on distortions for iterations of an interval map with a neutral fixed point was shown in the proof of [20, Lemma 5.3].

Lemma 4.8 (cf. [20, Lemma 5.3]).

Let r>0r>0 and let f:[0,r)f\colon[0,r)\to\mathbb{R} be a C2C^{2} map satisfying f0=0f0=0, f0=1f^{\prime}0=1 and fx>1f^{\prime}x>1 for all x(0,r)x\in(0,r). There exists a constant K>0K>0 such that for every nn\in\mathbb{N} and any pair x,yx,y of points in Jn1J_{n-1},

log|(fn)y||(fn)x|K|fnxfny|i=0n1|Ji||J0|,\log\frac{|(f^{n})^{\prime}y|}{|(f^{n})^{\prime}x|}\leq K|f^{n}x-f^{n}y|\sum_{i=0}^{n-1}\frac{|J_{i}|}{|J_{0}|},

where q0=rq_{0}=r, fqi+1=qifq_{i+1}=q_{i} and Ji=[qi+1,qi)J_{i}=[q_{i+1},q_{i}) for i=0,,n1i=0,\ldots,n-1.

We now proceed to distortion estimates of an induced potential. Notice that

Λ^=(ΛΔ(1))n=1Rn((1,0)).\widehat{\Lambda}=(\Lambda\setminus\varDelta(1))\setminus\bigcup_{n=1}^{\infty}R^{-n}((1^{\infty},0)).

Define an induced annealed geometric potential φ^:Λ^\widehat{\varphi}\colon\widehat{\Lambda}\to\mathbb{R} by

φ^(ω,x)=St(ω,x)φ(ω,x).\widehat{\varphi}(\omega,x)=S_{t(\omega,x)}\varphi(\omega,x).

For a pair (ω,x),(ϱ,y)(\omega,x),(\varrho,y) of distinct points in Λ^\widehat{\Lambda} contained in the same 11-cylinder, we introduce their separation time

s((ω,x),(ϱ,y))=min{n1:a1(R^n(ω,x))a1(R^n(ϱ,y))}.s((\omega,x),(\varrho,y))=\min\{n\geq 1\colon a_{1}(\widehat{R}^{n}(\omega,x))\neq a_{1}(\widehat{R}^{n}(\varrho,y))\}.

Note that s((ω,x),(ϱ,y))2s((\omega,x),(\varrho,y))\geq 2 implies t(ω,x)=t(ϱ,y)t(\omega,x)=t(\varrho,y). We evaluate the quantity

φ^(ω,x)φ^(ϱ,y)=log|(Tωt(ω,x))y||(Tωt(ω,x))x|.\widehat{\varphi}(\omega,x)-\widehat{\varphi}(\varrho,y)=\log\frac{|(T_{\omega}^{t(\omega,x)})^{\prime}y|}{|(T_{\omega}^{t(\omega,x)})^{\prime}x|}.
Lemma 4.9.

There exist constants K>0K>0 and τ(0,1)\tau\in(0,1) such that for any pair (ω,x),(ϱ,y)(\omega,x),(\varrho,y) of points in Λ^\widehat{\Lambda} with s((ω,x),(ϱ,y))2s((\omega,x),(\varrho,y))\geq 2,

φ^(ω,x)φ^(ϱ,y)Kτs((x,ω),(ϱ,y)).\widehat{\varphi}(\omega,x)-\widehat{\varphi}(\varrho,y)\leq K\tau^{s((x,\omega),(\varrho,y))}.
Proof.

For (ω,x),(ϱ,y)Λ^(\omega,x),(\varrho,y)\in\widehat{\Lambda} as in the statement, put

k=min{i1:Ri(ω,x)Δ(1)} and n=t(ω,x),k=\min\{i\geq 1\colon R^{i}(\omega,x)\in\varDelta(1)\}\text{ and }n=t(\omega,x),

and decompose Rn=RnkRkR^{n}=R^{n-k}\circ R^{k}. We estimate contributions from the first kk iteration and the remaining nkn-k iteration separately. Lemma 3.4 gives

(4.13) Skφ(ω,x)Skφ(ϱ,y)2|TωkxTϱky| if k=1.S_{k}\varphi(\omega,x)-S_{k}\varphi(\varrho,y)\leq 2|T^{k}_{\omega}x-T^{k}_{\varrho}y|\ \text{ if $k=1$}.

By Lemma 3.4 and Lemma 3.2,

(4.14) Skφ(ω,x)Skφ(ϱ,y)2i=1k|TωixTϱiy|2(1+i=1k1(49)(ki)/2)|TωkxTϱky| if k>1.\begin{split}S_{k}\varphi(\omega,x)-S_{k}\varphi(\varrho,y)&\leq 2\sum_{i=1}^{k}|T^{i}_{\omega}x-T^{i}_{\varrho}y|\\ &\leq 2\left(1+\sum_{i=1}^{k-1}\left(\frac{4}{9}\right)^{\lfloor(k-i)/2\rfloor}\right)|T^{k}_{\omega}x-T^{k}_{\varrho}y|\ \text{ if $k>1$}.\end{split}

Put τ=(4/9)14(0,1)\tau=(4/9)^{\frac{1}{4}}\in(0,1) and K0=2(1+i=0(4/9)i/2)K_{0}=2\left(1+\sum_{i=0}^{\infty}(4/9)^{\lfloor i/2\rfloor}\right). By the mean value theorem, there exists (θnω,z)Δ(an+1(ω,x))(\theta^{n}\omega,z)\in\varDelta(a_{n+1}(\omega,x)) such that

|TωkxTωky||TωnxTωny|=|Tωi=0s((ω,x),(ϱ,y))1t(R^i(ω,x))xTωi=0s((ω,x),(ϱ,y))1t(R^i(ω,x))y||(Tθnωi=0s((ω,x),(ϱ,y))1t(R^i(ω,x))n)z|.\begin{split}|T_{\omega}^{k}x-T_{\omega}^{k}y|&\leq|T_{\omega}^{n}x-T_{\omega}^{n}y|\\ &=\frac{|T_{\omega}^{\sum_{i=0}^{s((\omega,x),(\varrho,y))-1}t(\widehat{R}^{i}(\omega,x))}x-T_{\omega}^{\sum_{i=0}^{s((\omega,x),(\varrho,y))-1}t(\widehat{R}^{i}(\omega,x))}y|}{|(T_{\theta^{n}\omega}^{\sum_{i=0}^{s((\omega,x),(\varrho,y))-1}t(\widehat{R}^{i}(\omega,x))-n})^{\prime}z|}.\end{split}

By Lemma 3.2, there exists a uniform constant K1>0K_{1}>0 such that

(4.15) |TωnxTωny|1|(Tθωi=0s((ω,x),(ϱ,y))1t(R^i(ω,x))n)z|K1τ2s((ω,x),(ϱ,y)).|T_{\omega}^{n}x-T_{\omega}^{n}y|\leq\frac{1}{|(T_{\theta\omega}^{\sum_{i=0}^{s((\omega,x),(\varrho,y))-1}t(\widehat{R}^{i}(\omega,x))-n})^{\prime}z|}\leq K_{1}\tau^{2s((\omega,x),(\varrho,y))}.

By Lemma 4.7, there exists a uniform constant K2>0K_{2}>0 such that

(4.16) |Snkφ(Rk(ω,x))Snkφ(Rk(ϱ,y))|K2|TωnxTϱny|12.|S_{n-k}\varphi(R^{k}(\omega,x))-S_{n-k}\varphi(R^{k}(\varrho,y))|\leq K_{2}|T_{\omega}^{n}x-T_{\varrho}^{n}y|^{\frac{1}{2}}.

Combining (4.13), (4.14), (4.15) and (4.16) we obtain

φ^(ω,x)φ^(ϱ,y)=Snφ(ω,x)Snφ(ϱ,y)|Skφ(ω,x)Skφ(ϱ,y)|+|Snkφ(Rk(ω,x))Snkφ(Rk(ϱ,y))|K0K1τ2s((ω,x),(ϱ,y))+K2|TωnxTϱny|12(K0K1+K2K1)τs((ω,x),(ϱ,y)).\begin{split}\widehat{\varphi}(\omega,x)-\widehat{\varphi}(\varrho,y)&=S_{n}\varphi(\omega,x)-S_{n}\varphi(\varrho,y)\\ &\leq|S_{k}\varphi(\omega,x)-S_{k}\varphi(\varrho,y)|+|S_{n-k}\varphi(R^{k}(\omega,x))-S_{n-k}\varphi(R^{k}(\varrho,y))|\\ &\leq K_{0}K_{1}\tau^{2s((\omega,x),(\varrho,y))}+K_{2}|T_{\omega}^{n}x-T_{\varrho}^{n}y|^{\frac{1}{2}}\\ &\leq(K_{0}K_{1}+K_{2}\sqrt{K_{1}})\tau^{s((\omega,x),(\varrho,y))}.\end{split}

Setting K=K0K1+K2K1K=K_{0}K_{1}+K_{2}\sqrt{K_{1}} yields the desired inequality. ∎

For each nn\in\mathbb{N} define

Vn(φ^)=sup{φ^(ω,x)φ^(ϱ,y):(ω,x),(ϱ,y)Λ^,s((ω,x),(ϱ,y))n}.V_{n}(\widehat{\varphi})=\sup\{\widehat{\varphi}(\omega,x)-\widehat{\varphi}(\varrho,y)\colon(\omega,x),(\varrho,y)\in\widehat{\Lambda},\ s((\omega,x),(\varrho,y))\geq n\}.
Corollary 4.10.

There exist constants K>0K>0 and γ(0,1)\gamma\in(0,1) such that for every n1n\geq 1 we have Vn(φ^)KγnV_{n}(\widehat{\varphi})\leq K\gamma^{n}.

Proof.

Follows from Lemma 4.7 and Lemma 4.9.∎

4.6. Variational characterization of the annealed Gauss-Rényi measure

Define a potential ψ:\psi\colon\mathbb{N}^{\mathbb{N}}\to\mathbb{R} by

(4.17) ψ=φπ\psi=\varphi\circ\pi

and an induced potential ψ^:[1]\widehat{\psi}:\mathbb{N}^{\mathbb{N}}\setminus[1]\to\mathbb{R} by

(4.18) ψ^=φ^π|[1].\widehat{\psi}=\widehat{\varphi}\circ\pi|_{\mathbb{N}^{\mathbb{N}}\setminus[1]}.
Lemma 4.11.

The potential ψ\psi is unbounded and supψ<0\sup\psi<0. It is acceptable.

Proof.

The first assertion follows from the fact that φ\varphi is unbounded and supφ<0\sup\varphi<0. The second one follows from Rényi’s condition (3.1) and Lemma 3.3. ∎

The annealed Gauss-Rényi measure ηp\eta_{p} has the so-called ‘weak Gibbs property’.

Lemma 4.12.

There exists K1K\geq 1 such that for all n1n\geq 1, all a1anna_{1}\cdots a_{n}\in\mathbb{N}^{n} and all x[a1an]x\in[a_{1}\cdots a_{n}],

K1exp(Dn(φ))ηp([a1an]expSnψ(x)Kexp(Dn(φ)).K^{-1}\exp(-D_{n}(\varphi))\leq\frac{\eta_{p}([a_{1}\cdots a_{n}]}{\exp S_{n}\psi(x)}\leq K\exp(D_{n}(\varphi)).
Proof.

Follows from the fact that hph_{p} is bounded from above and away from 0.∎

Lemma 4.13.

We have P(ψ)=0P(\psi)=0.

Proof.

By Lemma 4.12, for all n1n\geq 1 and all a1anna_{1}\cdots a_{n}\in\mathbb{N}^{n} we have

K1exp(Dn(φ))ηp([a1an])sup[a1an]expSnψKexp(Dn(φ))ηp([a1an]).K^{-1}\exp(-D_{n}(\varphi))\eta_{p}([a_{1}\cdots a_{n}])\leq\sup_{[a_{1}\cdots a_{n}]}\exp S_{n}\psi\leq K\exp(D_{n}(\varphi))\eta_{p}([a_{1}\cdots a_{n}]).

Since ηp\eta_{p} is a probability and nn-cylinders are pairwise disjoint, summing the double inequalities over all a1anna_{1}\cdots a_{n}\in\mathbb{N}^{n}, taking logarithms, dividing by nn and using Lemma 3.5 we obtain P(ψ)=0P(\psi)=0. ∎

By Lemma 4.11 and Lemma 4.13, ψ\psi is acceptable and satisfies P(ψ)<P(\psi)<\infty. By Proposition 4.1, the variational principle holds for ψ\psi. Due to the existence of the neutral fixed point of the Rényi map T1T_{1}, ψ\psi is not locally Hölder continuous. Nevertheless the following holds.

Proposition 4.14.

The annealed Gauss-Rényi measure ηp\eta_{p} is the unique equilibrium state for the potential ψ\psi.

Proof.

A proof of Proposition 4.14 breaks into two steps. We first show that ηp\eta_{p} is an equilibrium state for the potential ψ\psi. We then establish the uniqueness of equilibrium state for the potential ψ\psi. To overcome the lack of regularity of ψ\psi in the second step, we take an inducing procedure that is now familiar in the construction of equilibrium states (see e.g., [27, Section 8], [30]).

Step 1: identifying ηp\eta_{p} as an equilibrium state. Since log|T0|\log|T_{0}^{\prime}| and log|T1|\log|T_{1}^{\prime}| are Lebesgue integrable, and since the Radon-Nikodým derivative hph_{p} is bounded from above, ψ\psi is ηp\eta_{p}-integrable. Since P(ψ)P(\psi) is finite by Lemma 4.13, the measure-theoretic entropy h(ηp)h(\eta_{p}) is finite (see §4.1). The family of 11-cylinders generates the Borel sigma algebra on \mathbb{N}^{\mathbb{N}}. Since hph_{p} is bounded from above and away from 0, using the Lebesgue measure on [0,1][0,1] and (3.2) one can show that kηp([k])logηp([k])-\sum_{k\in\mathbb{N}}\eta_{p}([k])\log\eta_{p}([k]) is finite. Since ηp\eta_{p} is mixing, it is ergodic. The Shannon-McMillan-Breimann theorem yields

limn1nlogηp([x1xn])=h(ηp) ηp-a.e.\lim_{n\to\infty}\frac{1}{n}\log\eta_{p}([x_{1}\cdots x_{n}])=-h(\eta_{p})\ \text{ $\eta_{p}$-a.e.}

Meanwhile, from Lemma 4.12 and Lemma 3.5 it follows that

limn1nlogηp([x1xn])=ψ𝑑ηp ηp-a.e.\lim_{n\to\infty}\frac{1}{n}\log\eta_{p}([x_{1}\cdots x_{n}])=\int\psi d\eta_{p}\ \text{ $\eta_{p}$-a.e.}

We have verified that h(ηp)+ψ𝑑ηp=0h(\eta_{p})+\int\psi d\eta_{p}=0. Since P(ψ)=0P(\psi)=0 by Lemma 4.13, ηp\eta_{p} is an equilibrium state for ψ\psi.

Step 2: establishing the uniquness of equilibrium state. Recall that σ^:^^\widehat{\sigma}\colon{\widehat{\mathbb{N}}}^{\mathbb{N}}\to{\widehat{\mathbb{N}}}^{\mathbb{N}} is the induced system associated with the inducing scheme ([1],t[1])(\mathbb{N}^{\mathbb{N}}\setminus[1],t_{\mathbb{N}^{\mathbb{N}}\setminus[1]}) of the left shift σ:\sigma\colon\mathbb{N}^{\mathbb{N}}\to\mathbb{N}^{\mathbb{N}} (see §4.2). For the induced potential ψ^\widehat{\psi} in (4.18), define Ψ:𝕄\Psi\colon\mathbb{M}^{\mathbb{N}}\to\mathbb{R} by

Ψ=ψ^ι.\Psi=\widehat{\psi}\circ\iota.
Lemma 4.15.

The potential Ψ\Psi is locally Hölder continuous.

Proof.

Follows from Corollary 4.10.∎

Next we compute the pressure P(Ψ)P(\Psi).

Lemma 4.16.

We have P(Ψ)=0P(\Psi)=0.

Proof.

Put K0=n=1varn(Ψ)K_{0}=\sum_{n=1}^{\infty}{\rm var}_{n}(\Psi). By Lemma 4.15, K0K_{0} is finite. For all n1n\geq 1 and all α1αn𝕄n\alpha_{1}\cdots\alpha_{n}\in\mathbb{M}^{n} we have

supη,ζ[α1αn](SnΨ(η)SnΨ(ζ))k=1nvark(Ψ)K0.\sup_{\eta,\zeta\in[\alpha_{1}\cdots\alpha_{n}]}\left(S_{n}\Psi(\eta)-S_{n}\Psi(\zeta)\right)\leq\sum_{k=1}^{n}{\rm var}_{k}(\Psi)\leq K_{0}.

Since hph_{p} is bounded from above and away from 0, there is a constant K11K_{1}\geq 1 such that for all n1n\geq 1 and all α1αn𝕄n\alpha_{1}\cdots\alpha_{n}\in\mathbb{M}^{n}, we have

K11ηp([α1αn])sup[α1αn]expSnΨK1ηp([α1αn]).K_{1}^{-1}\eta_{p}([\alpha_{1}\cdots\alpha_{n}])\leq\sup_{[\alpha_{1}\cdots\alpha_{n}]}\exp S_{n}\Psi\leq K_{1}\eta_{p}([\alpha_{1}\cdots\alpha_{n}]).

Summing these double inequalities over all α1αn𝕄n\alpha_{1}\cdots\alpha_{n}\in\mathbb{M}^{n},

K11α1αn𝕄nηp([α1αn])α1αn𝕄nsup[α1αn]expSnΨK1.K_{1}^{-1}\sum_{\alpha_{1}\cdots\alpha_{n}\in\mathbb{M}^{n}}\eta_{p}([\alpha_{1}\cdots\alpha_{n}])\leq\sum_{\alpha_{1}\cdots\alpha_{n}\in\mathbb{M}^{n}}\sup_{[\alpha_{1}\cdots\alpha_{n}]}\exp S_{n}\Psi\leq K_{1}.

By the definition of Λ^\widehat{\Lambda} and the fact that mpλpm_{p}\otimes\lambda_{p} has no atom,

α1αn𝕄nηp([α1αn])=ηp(Σ)=(mpλp)(Λ^)=(mpλp)(ΛΔ(1))>0.\sum_{\alpha_{1}\cdots\alpha_{n}\in\mathbb{M}^{n}}\eta_{p}([\alpha_{1}\cdots\alpha_{n}])=\eta_{p}(\Sigma)=(m_{p}\otimes\lambda_{p})(\widehat{\Lambda})=(m_{p}\otimes\lambda_{p})(\Lambda\setminus\varDelta(1))>0.

Hence, taking logarithms of the above double inequalities, dividing the result by nn and letting nn\to\infty yields P(Ψ)=0P(\Psi)=0. ∎

Since Ψ\Psi is acceptable by Lemma 4.15 and P(Ψ)P(\Psi) is finite by Lemma 4.16, the variational prinicple holds by Proposition 4.1. By Proposition 4.2 and P(Ψ)=0P(\Psi)=0 from Lemma 4.16, there exists a unique shift-invariant Gibbs state μ^(𝕄,σ)\widehat{\mu}\in\mathcal{M}(\mathbb{M}^{\mathbb{N}},\sigma), namely, there exists a constant K1K\geq 1 such that for every n1n\geq 1, every α1αn𝕄n\alpha_{1}\cdots\alpha_{n}\in\mathbb{M}^{n} and every z[α1αn]z\in[\alpha_{1}\cdots\alpha_{n}],

(4.19) K1μ^([α1αn])expSnΨ(z)K.K^{-1}\leq\frac{\widehat{\mu}([\alpha_{1}\cdots\alpha_{n}])}{\exp S_{n}\Psi(z)}\leq K.
Lemma 4.17.

Both t[1]ι𝑑μ^\int t_{\mathbb{N}^{\mathbb{N}}\setminus[1]}\circ\iota d\widehat{\mu} and Ψ𝑑μ^\int\Psi d\widehat{\mu} are finite.

Proof.

The function t[1]ιt_{\mathbb{N}^{\mathbb{N}}\setminus[1]}\circ\iota is constant on [α][\alpha] for each α𝕄\alpha\in\mathbb{M}. Let tαt_{\alpha} denote this constant. By the second inequality in (4.19), for all (ω,x)πι([α])(\omega,x)\in\pi\circ\iota([\alpha]) we have

μ^([α])K(1p)ptα1|(Tωtα)x|1K(1p)ptα1|Tωx|1.\widehat{\mu}([\alpha])\leq K(1-p)p^{t_{\alpha}-1}|(T_{\omega}^{t_{\alpha}})^{\prime}x|^{-1}\leq K(1-p)p^{t_{\alpha}-1}|T_{\omega}^{\prime}x|^{-1}.

For every k{1}k\in\mathbb{N}\setminus\{1\}, there is α𝕄\alpha\in\mathbb{M} such that π([α])Δ(k)\pi([\alpha])\subset\varDelta(k) and tα=nt_{\alpha}=n. Hence

(4.20) α𝕄tα=nμ^([α])K(1p)pn1(k=1supΔ(2k)|T0|1+k=2supΔ(2k1)|T1|1)2e2K(1p)pn1(k=1|J(2k)|+k=2|J(2k1)|)=3e2K(1p)pn1.\begin{split}\sum_{\begin{subarray}{c}\alpha\in\mathbb{M}\\ t_{\alpha}=n\end{subarray}}\widehat{\mu}([\alpha])&\leq K(1-p)p^{n-1}\left(\sum_{k=1}^{\infty}\sup_{\varDelta(2k)}|T_{0}^{\prime}|^{-1}+\sum_{k=2}^{\infty}\sup_{\varDelta(2k-1)}|T_{1}^{\prime}|^{-1}\right)\\ &\leq 2e^{2}K(1-p)p^{n-1}\left(\sum_{k=1}^{\infty}|J(2k)|+\sum_{k=2}^{\infty}|J(2k-1)|\right)\\ &=3e^{2}K(1-p)p^{n-1}.\end{split}

To deduce the second inequality we have used (3.1). Therefore

t[1]ι𝑑μ^=n=1nα𝕄tα=nμ^([α])<,\int t_{\mathbb{N}^{\mathbb{N}}\setminus[1]}\circ\iota d\widehat{\mu}=\sum_{n=1}^{\infty}n\sum_{\begin{subarray}{c}\alpha\in\mathbb{M}\\ t_{\alpha}=n\end{subarray}}\widehat{\mu}([\alpha])<\infty,

as required.

There exist constants K>0K>0 and c>1c>1 such that if nn\in\mathbb{N} and xJ(1)x\in J(1) are such that x,,T1n1xJ(1)x,\ldots,T_{1}^{n-1}x\in J(1) then |(T1n)x|Kcn|(T_{1}^{n})^{\prime}x|\leq Kc^{n}. Moreover, cc can be taken arbitrarily close to 11 at the expense of enlarging KK. Now, let nn\in\mathbb{N}, α𝕄\alpha\in\mathbb{M} satisfy tα=nt_{\alpha}=n. For ζ=(ω,x)[α]\zeta=(\omega,x)\in[\alpha] we have

Ψ(ζ)=logp(ω1)log|(Tω1)x|+(n1)logplog|(T1n1)Tω1x|,\Psi(\zeta)=\log p(\omega_{1})-\log|(T_{\omega_{1}})^{\prime}x|+(n-1)\log p-\log|(T_{1}^{n-1})^{\prime}T_{\omega_{1}}x|,

where Tω1x,,Tω1n1xJ(1)T_{\omega_{1}}x,\ldots,T_{\omega_{1}}^{n-1}x\in J(1) provided n2n\geq 2. It follows that there exists a constant K>0K>0 independent of nn, α\alpha, ζ\zeta such that

(4.21) |Ψ(ζ)|Kn.|\Psi(\zeta)|\leq Kn.

From (4.20) and (4.21) we obtain

|Ψ𝑑μ^||Ψ|𝑑μ^n=1α𝕄tα=nμ^([α])sup[α]|Ψ|n=1Knα𝕄tα=nμ^([α])<,\left|\int\Psi d\widehat{\mu}\right|\leq\int|\Psi|d\widehat{\mu}\leq\sum_{n=1}^{\infty}\sum_{\begin{subarray}{c}\alpha\in\mathbb{M}\\ t_{\alpha}=n\end{subarray}}\widehat{\mu}([\alpha])\sup_{[\alpha]}|\Psi|\leq\sum_{n=1}^{\infty}Kn\sum_{\begin{subarray}{c}\alpha\in\mathbb{M}\\ t_{\alpha}=n\end{subarray}}\widehat{\mu}([\alpha])<\infty,

as required. ∎

Since Ψ𝑑μ^\int\Psi d\widehat{\mu} is finite by Lemma 4.17, μ^\widehat{\mu} is the unique equilibrium state for the potential Ψ\Psi by Proposition 4.2. In particular we have

(4.22) P(Ψ)=h(μ^)+Ψ𝑑μ^.P(\Psi)=h(\widehat{\mu})+\int\Psi d\widehat{\mu}.

By the finiteness of t[1]ι𝑑μ^\int t_{\mathbb{N}^{\mathbb{N}}\setminus[1]}\circ\iota d\widehat{\mu} in Lemma 4.17, the measure

μ=1t[1]ι𝑑μ^n=1i=0n1μ^|{t[1]ι=n}ι1σi\mu=\frac{1}{\int t_{\mathbb{N}^{\mathbb{N}}\setminus[1]}\circ\iota d\widehat{\mu}}\sum_{n=1}^{\infty}\sum_{i=0}^{n-1}\widehat{\mu}|_{\{t_{\mathbb{N}^{\mathbb{N}}\setminus[1]}\circ\iota=n\}}\circ\iota^{-1}\circ\sigma^{-i}

belongs to (,σ),\mathcal{M}(\mathbb{N}^{\mathbb{N}},\sigma), and by Abramov-Kac’s formula [30, Theorem 2.3]

(4.23) h(μ^)+Ψ𝑑μ^=(h(μ)+ψ𝑑μ)t[1]ι𝑑μ^.h(\widehat{\mu})+\int\Psi d\widehat{\mu}=\left(h(\mu)+\int\psi d\mu\right)\int t_{\mathbb{N}^{\mathbb{N}}\setminus[1]}\circ\iota d\widehat{\mu}.

Combining (4.22), (4.23) and P(Ψ)=0P(\Psi)=0 in Lemma 4.16 we obtain h(μ)+ψ𝑑μ=0h(\mu)+\int\psi d\mu=0. Since P(ψ)=0P(\psi)=0 by Lemma 4.13, μ\mu is an equilibrium state for the potential ψ\psi.

We claim that μ\mu is the unique equilibrium state for the potential ψ\psi. Indeed, let νψ(,σ)\nu\in\mathcal{M}_{\psi}(\mathbb{N}^{\mathbb{N}},\sigma) be an equilibrium state for ψ\psi with ν(^)>0\nu(\widehat{\mathbb{N}}^{\mathbb{N}})>0. The normalized restriction of ν\nu to ^\widehat{\mathbb{N}}^{\mathbb{N}}, denoted by ν^\widehat{\nu}, belongs to (^,σ^^)\mathcal{M}(\widehat{\mathbb{N}}^{\mathbb{N}},\widehat{\sigma}_{\widehat{\mathbb{N}}^{\mathbb{N}}}). From P(ψ)=0P(\psi)=0, Abramov-Kac’s formula and P(Ψ)=0P(\Psi)=0, ν^\widehat{\nu} is an equilibrium state for the potential Ψ\Psi, namely μ^=ν^\widehat{\mu}=\widehat{\nu}. It follows that μ=ν\mu=\nu. Moreover, the only measure in ψ(,σ)\mathcal{M}_{\psi}(\mathbb{N}^{\mathbb{N}},\sigma) which does not give positive weight to ^\widehat{\mathbb{N}}^{\mathbb{N}} is the unit point mass at π1(1,0)\pi^{-1}(1^{\infty},0), which is precisely the fixed point of σ\sigma in the 11-cylinder [1][1]. Since h(δπ1(1,0))=0h(\delta_{\pi^{-1}(1^{\infty},0)})=0 and |T10|=1|T_{1}^{\prime}0|=1, we have h(δπ1(1,0))+ψ𝑑δπ1(1,0)=logp<0=P(ψ).h(\delta_{\pi^{-1}(1^{\infty},0)})+\int\psi d\delta_{\pi^{-1}(1^{\infty},0)}=\log p<0=P(\psi). Therefore the claim holds. The proof of Proposition 4.14 is complete. ∎

4.7. Proof of Theorem 1.4

We define a sequence (ν~n)n=1(\tilde{\nu}_{n})_{n=1}^{\infty} of Borel probability measures on ()\mathcal{M}(\mathbb{N}^{\mathbb{N}}) replacing ϕ\phi in (4.4) by ψ\psi in (4.17). Define a parametrized family of twisted induced potentials Ψγ:𝕄\Psi_{\gamma}\colon\mathbb{M}^{\mathbb{N}}\to\mathbb{R} (γ)(\gamma\in\mathbb{R}) replacing ϕ\phi in (4.6) by ψ\psi. Then Ψγ\Psi_{\gamma} is locally Hölder continuous for all γ\gamma\in\mathbb{R} by Lemma 4.15, and P(Ψ0)=0P(\Psi_{0})=0 by Lemma 4.16. By Theorem 4.4, (ν~n)n=1(\tilde{\nu}_{n})_{n=1}^{\infty} is exponentially tight and satisfies the LDP with the good rate function IψI_{\psi}.

The coding map π:Λ\pi\colon\mathbb{N}^{\mathbb{N}}\to\Lambda in (4.7) induces a continuous map π:ν()νπ1(Λ)\pi_{*}\colon\nu\in\mathcal{M}(\mathbb{N}^{\mathbb{N}})\mapsto\nu\circ\pi^{-1}\in\mathcal{M}(\Lambda). Since ν~nπ1=μ~n\tilde{\nu}_{n}\circ\pi_{*}^{-1}=\tilde{\mu}_{n} for every n1n\geq 1, by the Contraction Principle in Proposition 2.3, (μ~n)n=1(\tilde{\mu}_{n})_{n=1}^{\infty} is exponentially tight and satisfies the LDP with the good rate function IpI_{p} given by

Ip(μ)=inf{Iψ(ν):ν(),π(ν)=μ}.I_{p}(\mu)=\inf\{I_{\psi}(\nu)\colon\nu\in\mathcal{M}(\mathbb{N}^{\mathbb{N}}),\ \pi_{*}(\nu)=\mu\}.

Since IψI_{\psi} is convex, so is IpI_{p}. Since ηp\eta_{p} is an equilibrium state for ψ\psi by Proposition 4.14, it is a minimizer of IψI_{\psi}. The equation π(ηp)=mpλp\pi_{*}(\eta_{p})=m_{p}\otimes\lambda_{p} shows that mpλpm_{p}\otimes\lambda_{p} is a minimizer of IpI_{p}.

By the last assertion of Proposition 2.3, to conclude the uniqueness of minimizer of IpI_{p} it suffices to show the uniqueness of minimizer of IψI_{\psi}. Since ψ\psi is acceptable by Lemma 4.11, it is uniformly continuous. By virtue of Proposition 4.5, it suffices to show β(ψ)<1\beta_{\infty}(\psi)<1. Direct calculations show that there exist constants K1>K0>0K_{1}>K_{0}>0 such that

4K0(1p)k(k+2)sup[k]eψ4K1(1p)k(k+2)\frac{4K_{0}(1-p)}{k(k+2)}\leq\sup_{[k]}e^{\psi}\leq\frac{4K_{1}(1-p)}{k(k+2)}

for all k0k\in\mathbb{N}_{0}, and

4K0p(k+1)(k+3)sup[k]eψ4K1p(k+1)(k+3)\frac{4K_{0}p}{(k+1)(k+3)}\leq\sup_{[k]}e^{\psi}\leq\frac{4K_{1}p}{(k+1)(k+3)}

for all k1k\in\mathbb{N}_{1}. Since sup[k]eβψ=(sup[k]eψ)β\sup_{[k]}e^{\beta\psi}=(\sup_{[k]}e^{\psi})^{\beta}, these estimates imply β(ψ)=1/2\beta_{\infty}(\psi)=1/2.

The deduction of Theorem 1.4(b) from Theorem 1.4(a) is much simpler than that of Theorem 1.5(b) from Theorem 1.5(a) carried out in §3.5. The exponential tightness in Theorem 1.4(a) implies the tightness, which ensures the existence of a limit point by Prohorov’s theorem. The LDP and the uniqueness of minimizer in Theorem 1.4(a) together rule out the existence of a limit point that is different from the unit point mass at the minimizer. The proof of Theorem 1.4 is complete. ∎

4.8. Annealed and quenched level-1 large deviations for the Gauss-Rényi map

For p(0,1)p\in(0,1) and a bounded continuous function f:Λf\colon\Lambda\to\mathbb{R}, define a function Ip,f:[0,]I_{p,f}\colon\mathbb{R}\to[0,\infty] by

Ip,f(α)=inf{Ip(ν):ν(Λ),f𝑑ν=α}.I_{p,f}(\alpha)=\inf\left\{I_{p}(\nu)\colon\nu\in\mathcal{M}(\Lambda),\ \int fd\nu=\alpha\right\}.

By Theorem 1.4(a), Ip,fI_{p,f} is convex and vanishes only at the mean α=fd(mpλp)\alpha=\int fd(m_{p}\otimes\lambda_{p}). Put

f¯=inf{f𝑑ν:ν(Λ)} and f¯=sup{f𝑑ν:ν(Λ)}.\underline{f}=\inf\left\{\int fd\nu\colon\nu\in\mathcal{M}(\Lambda)\right\}\ \text{ and }\ \overline{f}=\sup\left\{\int fd\nu\colon\nu\in\mathcal{M}(\Lambda)\right\}.

The next corollary of independent interest follows from the Contraction Principle applied to the level-2 LDP in Theorem 1.4(a).

Corollary 4.18 (annealed level-1 LDP).

Let f:Λf\colon\Lambda\to\mathbb{R} be a bounded continuous function such that f¯<f¯\underline{f}<\overline{f}. For any p(0,1)p\in(0,1) the following statements hold:

  • (a)

    if fd(mpλp)<αf¯\int fd(m_{p}\otimes\lambda_{p})<\alpha\leq\overline{f} then

    limn1nlog(ω,x)Fix(Rn)(1/n)i=0n1f(Ri(ω,x))αQpn(ω)|(Tωn)x|1=Ip,f(α)<0;\lim_{n\to\infty}\frac{1}{n}\log\sum_{\begin{subarray}{c}(\omega,x)\in{\rm Fix}(R^{n})\\ (1/n)\sum_{i=0}^{n-1}f(R^{i}(\omega,x))\geq\alpha\end{subarray}}Q_{p}^{n}(\omega)|(T^{n}_{\omega})^{\prime}x|^{-1}=-I_{p,f}(\alpha)<0;
  • (b)

    if f¯α<fd(mpλp)\underline{f}\leq\alpha<\int fd(m_{p}\otimes\lambda_{p}) then

    limn1nlog(ω,x)Fix(Rn)(1/n)k=0n1f(Rk(ω,x))αQpn(ω)|(Tωn)x|1=Ip,f(α)<0.\lim_{n\to\infty}\frac{1}{n}\log\sum_{\begin{subarray}{c}(\omega,x)\in{\rm Fix}(R^{n})\\ (1/n)\sum_{k=0}^{n-1}f(R^{k}(\omega,x))\leq\alpha\end{subarray}}Q_{p}^{n}(\omega)|(T^{n}_{\omega})^{\prime}x|^{-1}=-I_{p,f}(\alpha)<0.

We apply Corollary 4.18 to the problem of frequency of digits in the random continued fraction expansion (1.1). Recall the algorithm in §2.1, and let us use the square bracket to denote the 22-cylinders in Ω\Omega: for i,j{0,1}i,j\in\{0,1\},

[ij]={ωΩ:ω1=i,ω2=j}.[ij]=\{\omega\in\Omega\colon\omega_{1}=i,\omega_{2}=j\}.

Let nn\in\mathbb{N} and (ω,x)Λ(\omega,x)\in\Lambda. For each kk\in\mathbb{N}, Cn(ω,x)=kC_{n}(\omega,x)=k holds if and only if C(Rn1(ω,x))=kC(R^{n-1}(\omega,x))=k and ωn+1=0\omega_{n+1}=0, or else C(Rn1(ω,x))=k1C(R^{n-1}(\omega,x))=k-1 and ωn+1=1\omega_{n+1}=1. For each mm\in\mathbb{N}, C(ω,x)=mC(\omega,x)=m holds if and only if 1/x=m\lfloor 1/x\rfloor=m and ω1=0\omega_{1}=0, or else 1/(1x)=m\lfloor 1/(1-x)\rfloor=m and ω1=1\omega_{1}=1.

If k=1k=1 then define

Ak=[00]×(1k+1,1k].A_{k}=[00]\times\left(\frac{1}{k+1},\frac{1}{k}\right].

If k2k\geq 2 then define

Ak=([00]×(1k+1,1k])([10]×[k1k,kk+1))([01]×(1k,1k1])([11]×[k2k1,k1k)).\begin{split}A_{k}=&\left([00]\times\left(\frac{1}{k+1},\frac{1}{k}\right]\right)\cup\left([10]\times\left[\frac{k-1}{k},\frac{k}{k+1}\right)\right)\\ &\cup\left([01]\times\left(\frac{1}{k},\frac{1}{k-1}\right]\right)\cup\left([11]\times\left[\frac{k-2}{k-1},\frac{k-1}{k}\right)\right).\end{split}

Notice that Cn(ω,x)=kC_{n}(\omega,x)=k holds if and only if Rn1(ω,x)AkR^{n-1}(\omega,x)\in A_{k}. Let 1lk:Λ\mbox{1}\hskip-2.5pt\mbox{l}_{k}\colon\Lambda\to\mathbb{R} denote the indicator function of AkΛA_{k}\cap\Lambda. Let p(0,1)p\in(0,1). By Birkhoff’s ergodic theorem, for mpλpm_{p}\otimes\lambda_{p}-almost every (ω,x)Λ(\omega,x)\in\Lambda we have

limn#{1in:Ci(ω,x)=k}n=1lkd(mpλp).\lim_{n\to\infty}\frac{\#\{1\leq i\leq n\colon C_{i}(\omega,x)=k\}}{n}=\int\mbox{1}\hskip-2.5pt\mbox{l}_{k}d(m_{p}\otimes\lambda_{p}).

Clearly, 1lk\mbox{1}\hskip-2.5pt\mbox{l}_{k} is bounded continuous and satisfies 1lk¯=0\underline{\mbox{1}\hskip-2.5pt\mbox{l}_{k}}=0, 1lk¯=1\overline{\mbox{1}\hskip-2.5pt\mbox{l}_{k}}=1, 0<1lkd(mpλp)<10<\int\mbox{1}\hskip-2.5pt\mbox{l}_{k}d(m_{p}\otimes\lambda_{p})<1. By Corollary 4.18 the following hold:

  • if 1lkd(mpλp)<α1\int\mbox{1}\hskip-2.5pt\mbox{l}_{k}d(m_{p}\otimes\lambda_{p})<\alpha\leq 1 then

    limn1nlog(ω,x)Fix(Rn)#{1in:Ci(ω,x)=k}nαQpn(ω)|(Tωn)x|1=Ip,1lk(α)<0;\lim_{n\to\infty}\frac{1}{n}\log\sum_{\begin{subarray}{c}(\omega,x)\in{\rm Fix}(R^{n})\\ \frac{\#\{1\leq i\leq n\colon C_{i}(\omega,x)=k\}}{n}\geq\alpha\end{subarray}}Q_{p}^{n}(\omega)|(T^{n}_{\omega})^{\prime}x|^{-1}=-I_{p,\mbox{1}\hskip-2.04861pt\mbox{l}_{k}}(\alpha)<0;
  • if 0α<1lkd(mpλp)0\leq\alpha<\int\mbox{1}\hskip-2.5pt\mbox{l}_{k}d(m_{p}\otimes\lambda_{p}) then

    limn1nlog(ω,x)Fix(Rn)#{1in:Ci(ω,x)=k}nαQpn(ω)|(Tωn)x|1=Ip,1lk(α)<0.\lim_{n\to\infty}\frac{1}{n}\log\sum_{\begin{subarray}{c}(\omega,x)\in{\rm Fix}(R^{n})\\ \frac{\#\{1\leq i\leq n\colon C_{i}(\omega,x)=k\}}{n}\leq\alpha\end{subarray}}Q_{p}^{n}(\omega)|(T^{n}_{\omega})^{\prime}x|^{-1}=-I_{p,\mbox{1}\hskip-2.04861pt\mbox{l}_{k}}(\alpha)<0.

Recall the notation in §3.2. If n2n\geq 2 then the indicator function of AkA_{k} is constant on each nn-cylinder Δ(a1an)\varDelta(a_{1}\cdots a_{n}). Moreover, each nn-cylinder contains exactly one point from Fix(Rn){\rm Fix}(R^{n}), and if (ω,x)Δ(a1an)Fix(Rn)(\omega,x)\in\varDelta(a_{1}\cdots a_{n})\cap{\rm Fix}(R^{n}) then by Lemma 3.5, Qpn(ω)|(Tωn)x|1Q_{p}^{n}(\omega)|(T^{n}_{\omega})^{\prime}x|^{-1} is comparable to (mpλp)(Δ(a1an))(m_{p}\otimes\lambda_{p})(\varDelta(a_{1}\cdots a_{n})) up to the subexponential factor exp(Dn(φ))\exp(D_{n}(\varphi)). Hence, the above annealed level-1 LDP for periodic points of RR extends to an annealed level-1 LDP for mpλpm_{p}\otimes\lambda_{p}-typical points:

  • if 1lkd(mpλp)<α1\int\mbox{1}\hskip-2.5pt\mbox{l}_{k}d(m_{p}\otimes\lambda_{p})<\alpha\leq 1 then

    limn1nlog(mpλp){(ω,x)Λ:#{1in:Ci(ω,x)=k}nα}=Ip,1lk(α);\lim_{n\to\infty}\frac{1}{n}\log(m_{p}\otimes\lambda_{p})\left\{(\omega,x)\in\Lambda\colon\frac{\#\{1\leq i\leq n\colon C_{i}(\omega,x)=k\}}{n}\geq\alpha\right\}=-I_{p,\mbox{1}\hskip-2.04861pt\mbox{l}_{k}}(\alpha);
  • if 0α<1lkd(mpλp)0\leq\alpha<\int\mbox{1}\hskip-2.5pt\mbox{l}_{k}d(m_{p}\otimes\lambda_{p}) then

    limn1nlog(mpλp){(ω,x)Λ:#{1in:Ci(ω,x)=k}nα}=Ip,1lk(α).\lim_{n\to\infty}\frac{1}{n}\log(m_{p}\otimes\lambda_{p})\left\{(\omega,x)\in\Lambda\colon\frac{\#\{1\leq i\leq n\colon C_{i}(\omega,x)=k\}}{n}\leq\alpha\right\}=-I_{p,\mbox{1}\hskip-2.04861pt\mbox{l}_{k}}(\alpha).

We now move on to a quenched counterpart. The next corollary of independent interest is a consequence of Theorem 1.5(a). Since it only gives an upper bound for closed sets, we only get inequalities for upper limits which should not be optimal.

Corollary 4.19 (quenched level-1 upper bounds).

Let f:Λf\colon\Lambda\to\mathbb{R} be a bounded continuous function such that f¯<f¯\underline{f}<\overline{f}. For any p(0,1)p\in(0,1) the following statements hold:

  • (a)

    if fd(mpλp)<αf¯\int fd(m_{p}\otimes\lambda_{p})<\alpha\leq\overline{f} then for mpm_{p}-almost every ωΩ\omega\in\Omega,

    lim supn1nlogxFix(Tωn)(1/n)i=0n1f(Tωix)α|(Tωn)x|1Ip,f(α)<0;\limsup_{n\to\infty}\frac{1}{n}\log\sum_{\begin{subarray}{c}x\in{\rm Fix}(T_{\omega}^{n})\\ (1/n)\sum_{i=0}^{n-1}f(T_{\omega}^{i}x)\geq\alpha\end{subarray}}|(T^{n}_{\omega})^{\prime}x|^{-1}\leq-I_{p,f}(\alpha)<0;
  • (b)

    if f¯α<fd(mpλp)\underline{f}\leq\alpha<\int fd(m_{p}\otimes\lambda_{p}) then for mpm_{p}-almost every ωΩ\omega\in\Omega,

    lim supn1nlogxFix(Tωn)(1/n)i=0n1f(Tωix)α|(Tωn)x|1Ip,f(α)<0.\limsup_{n\to\infty}\frac{1}{n}\log\sum_{\begin{subarray}{c}x\in{\rm Fix}(T_{\omega}^{n})\\ (1/n)\sum_{i=0}^{n-1}f(T_{\omega}^{i}x)\leq\alpha\end{subarray}}|(T^{n}_{\omega})^{\prime}x|^{-1}\leq-I_{p,f}(\alpha)<0.

Let p(0,1)p\in(0,1) and kk\in\mathbb{N}. By Birkhoff’s ergodic theorem and Fubini’s theorem, for mpm_{p}-almost every ωΩ\omega\in\Omega and λp\lambda_{p}-almost every xΛωx\in\Lambda_{\omega} we have

limn#{1in:Ci(ω,x)=k}n=1lkd(mpλp).\lim_{n\to\infty}\frac{\#\{1\leq i\leq n\colon C_{i}(\omega,x)=k\}}{n}=\int\mbox{1}\hskip-2.5pt\mbox{l}_{k}d(m_{p}\otimes\lambda_{p}).

Corollary 4.19 yields the following:

  • if 1lkd(mpλp)<α1\int\mbox{1}\hskip-2.5pt\mbox{l}_{k}d(m_{p}\otimes\lambda_{p})<\alpha\leq 1 then for mpm_{p}-almost every ωΩ\omega\in\Omega,

    lim supn1nlogxFix(Tωn)#{1in:Ci(ω,x)=k}nα|(Tωn)x|1Ip,1lk(α);\limsup_{n\to\infty}\frac{1}{n}\log\sum_{\begin{subarray}{c}x\in{\rm Fix}(T_{\omega}^{n})\\ \frac{\#\{1\leq i\leq n\colon C_{i}(\omega,x)=k\}}{n}\geq\alpha\end{subarray}}|(T^{n}_{\omega})^{\prime}x|^{-1}\leq-I_{p,\mbox{1}\hskip-2.04861pt\mbox{l}_{k}}(\alpha);
  • if 0α<1lkd(mpλp)0\leq\alpha<\int\mbox{1}\hskip-2.5pt\mbox{l}_{k}d(m_{p}\otimes\lambda_{p}) then for mpm_{p}-almost every ωΩ\omega\in\Omega,

    lim supn1nlogxFix(Tωn)#{1in:Ci(ω,x)=k}nα|(Tωn)x|1Ip,1lk(α).\limsup_{n\to\infty}\frac{1}{n}\log\sum_{\begin{subarray}{c}x\in{\rm Fix}(T_{\omega}^{n})\\ \frac{\#\{1\leq i\leq n\colon C_{i}(\omega,x)=k\}}{n}\leq\alpha\end{subarray}}|(T^{n}_{\omega})^{\prime}x|^{-1}\leq-I_{p,\mbox{1}\hskip-2.04861pt\mbox{l}_{k}}(\alpha).

Recall the notation in §3.2 again. Let ωΩ\omega\in\Omega, nn\in\mathbb{N} and let a1ana_{1}\cdots a_{n}\in\mathbb{N}^{\mathbb{N}} satisfy ωiai\omega_{i}\equiv a_{i} mod 22 for i=1,,ni=1,\ldots,n. If n2n\geq 2 then the restriction of the indicator function of AkA_{k} to {ω}×J(a1an)\{\omega\}\times J(a_{1}\cdots a_{n}) is constant. Clearly, J(a1an)Fix(Tωn)J(a_{1}\cdots a_{n})\cap{\rm Fix}(T_{\omega}^{n}) is a singleton. If xJ(a1an)Fix(Tωn)x\in J(a_{1}\cdots a_{n})\cap{\rm Fix}(T_{\omega}^{n}), then by Lemma 3.5, |(Tωn)x|1|(T^{n}_{\omega})^{\prime}x|^{-1} is comparable to λp(J(a1an))\lambda_{p}(J(a_{1}\cdots a_{n})) up to the subexponential factor exp(Dn(φ))\exp(D_{n}(\varphi)). Hence, the above quenched level-1 upper bounds extend to quenched level-1 upper bounds for λp\lambda_{p}-typical points:

  • if 1lkd(mpλp)<α1\int\mbox{1}\hskip-2.5pt\mbox{l}_{k}d(m_{p}\otimes\lambda_{p})<\alpha\leq 1 then for mpm_{p}-almost every ωΩ\omega\in\Omega,

    lim supn1nlogλp{x(0,1):#{1in:Ci(ω,x)=k}nα}Ip,1lk(α);\limsup_{n\to\infty}\frac{1}{n}\log\lambda_{p}\left\{x\in(0,1)\setminus\mathbb{Q}\colon\frac{\#\{1\leq i\leq n\colon C_{i}(\omega,x)=k\}}{n}\geq\alpha\right\}\leq-I_{p,\mbox{1}\hskip-2.04861pt\mbox{l}_{k}}(\alpha);
  • if 0α<1lkd(mpλp)0\leq\alpha<\int\mbox{1}\hskip-2.5pt\mbox{l}_{k}d(m_{p}\otimes\lambda_{p}) then for mpm_{p}-almost every ωΩ\omega\in\Omega,

    lim supn1nlogλp{x(0,1):#{1in:Ci(ω,x)=k}nα}Ip,1lk(α).\limsup_{n\to\infty}\frac{1}{n}\log\lambda_{p}\left\{x\in(0,1)\setminus\mathbb{Q}\colon\frac{\#\{1\leq i\leq n\colon C_{i}(\omega,x)=k\}}{n}\leq\alpha\right\}\leq-I_{p,\mbox{1}\hskip-2.04861pt\mbox{l}_{k}}(\alpha).

Appendix A Periodic continued fractions

The classical Lagrange theorem asserts that the regular continued fraction expansion of a quadratic irrational is eventually periodic. So, any quadratic irrational in (0,1)(0,1) is eventually periodic under the iteration of the Gauss map. This appendix is a brief summary of known characterizations of periodic continued fractions in terms of iterations of the Gauss and Rényi maps. For a quadratic irrational xx\in\mathbb{R}, let xx^{\dagger} denote its Galois conjugate.

Proposition A.1 ([16]).

Let x(0,1)x\in(0,1). The following are equivalent:

  • (a)

    xx is a quadratic irrational and x<1x^{\dagger}<-1.

  • (b)

    There exists nn\in\mathbb{N} such that T0nx=xT_{0}^{n}x=x.

Although much less known, statements analogous to Proposition A.1 hold for the Rényi map.

Proposition A.2.

Let x(0,1)x\in(0,1). The following are equivalent:

  • (a)

    xx is a quadratic irrational and x<0x^{\dagger}<0.

  • (b)

    There exists nn\in\mathbb{N} such that T1nx=xT_{1}^{n}x=x.

For the reader’s convenience we include a proof of Proposition A.2 below. The idea is to translate analogous statements in [22] on the minus continued fraction to the backward continued fraction via simple algebraic manipulations.

Let xx\in\mathbb{R}. We define a sequence (xn)n=0(x_{n})_{n=0}^{\infty} of real numbers by

x0=x and xn=1xn1+1xn1 for n1.x_{0}=x\ \text{ and }\ x_{n}=\frac{1}{\lfloor x_{n-1}\rfloor+1-x_{n-1}}\ \text{ for }n\geq 1.

For n0n\geq 0 put

Dn(x)=xn+1.D_{n}(x)=\lfloor x_{n}\rfloor+1.

For n1n\geq 1, note that Dn(x)2D_{n}(x)\geq 2 since xn1x_{n}\geq 1. For n1n\geq 1 we set

rn(x)=D0(x)1  D1(x)1  Dn(x).r_{n}(x)=D_{0}(x)-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{D_{1}(x)}\hfill}}-\cdots-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{D_{n}(x)}\hfill}}.

By [22, Theorem 1.1] we obtain x=limnrn(x)x=\lim_{n}r_{n}(x), which is the minus continued fraction expansion of xx:

x=D0(x)1  D1(x)1  D2(x)1  Dn(x).x=D_{0}(x)-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{D_{1}(x)}\hfill}}-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{D_{2}(x)}\hfill}}-\cdots-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{D_{n}(x)}\hfill}}-\cdots.

We say xx has a purely periodic minus continued fraction expansion of period N+1N+1 if there exists NN\in\mathbb{N} such that

x=D0(x)1  D1(x)1  D2(x)1  DN(x)1  x.x=D_{0}(x)-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{D_{1}(x)}\hfill}}-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{D_{2}(x)}\hfill}}-\cdots-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{D_{N}(x)}\hfill}}-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{x}\hfill}}.
Proposition A.3 ([22, Theorem 1.4]).

Let xx\in\mathbb{R} be a quadratic irrational. Then xx has a purely periodic minus continued fraction expansion if and only if x>1x>1 and 0<x<10<x^{\dagger}<1.

Proof of Proposition A.2.

Let x(0,1)x\in(0,1) be a quadratic irrational. There is a quadratic equation az2+bz+c=0az^{2}+bz+c=0 with integer coefficients whose solutions are x,xx,x^{\dagger}. This equation is equivalent to a(1z)2(b+2a)(1z)+(a+b+c)=0a(1-z)^{2}-(b+2a)(1-z)+(a+b+c)=0. We have a+b+c0a+b+c\neq 0, for otherwise z=1z=1 would be a solution of the equation. For z{x,x}z\in\{x,x^{\dagger}\} we have

(a+b+c)((1z)1)2(b+2a)(1z)1+a=0.(a+b+c)\Bigl{(}(1-z)^{-1}\Bigr{)}^{2}-(b+2a)(1-z)^{-1}+a=0.

Hence, (1x)1(1-x)^{-1} is a quadratic irrational whose Galois conjugate is (1x)1(1-x^{\dagger})^{-1}.

Let x(0,1)x\in(0,1) be a quadratic irrational and suppose x<0x^{\dagger}<0. Then 0<(1x)1<10<(1-x^{\dagger})^{-1}<1 holds. Since (1x)1>1(1-x)^{-1}>1, by Proposition A.3 there exists an integer n2n\geq 2 such that the minus continued fraction expansion of (1x)1(1-x)^{-1} is periodic of period of nn:

11x=D0(x)1  D1(x)1  Dn1(x)1  D0(x)1  Dn1(x),\frac{1}{1-x}=D_{0}(x)-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{D_{1}(x)}\hfill}}-\cdots-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{D_{n-1}(x)}\hfill}}-\cdots\-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{D_{0}(x)}\hfill}}-\cdots-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{D_{n-1}(x)}\hfill}}-\cdots,

where Di(x)2D_{i}(x)\geq 2 for i=0,,n1i=0,\ldots,n-1. Rearranging this equality gives

x=11  D0(x)1  Dn1(x)1  D0(x).x=1-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{D_{0}(x)}\hfill}}-\cdots-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{D_{n-1}(x)}\hfill}}-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{D_{0}(x)}\hfill}}-\cdots.

From this and the uniqueness of the backward continued fraction given by the Rényi map T1T_{1}, we obtain T1nx=xT^{n}_{1}x=x.

Conversely, suppose there exists nn\in\mathbb{N} such that T1nx=xT_{1}^{n}x=x. Then the backward continued fraction of xx given by T1T_{1} is periodic of period nn, and we have

x=11  B1(x)1  Bn(x)1x,x=1-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{B_{1}(x)}\hfill}}-\cdots-\frac{\displaystyle{\hfill{1}\hfill\;\vrule}}{\displaystyle{\vrule\;\hfill{B_{n}(x)-1-x}\hfill}},

where Bi(x)=1/(1T1i1x)+1B_{i}(x)=\lfloor 1/(1-T_{1}^{i-1}x)\rfloor+1 for i=1,,ni=1,\ldots,n. Since this fraction can be represented by ax+b/(cx+d)ax+b/(cx+d) for some a,b,c,da,b,c,d\in\mathbb{Z} with adbc=1ad-bc=1 (see e.g., [19]), xx is a quadratic irrational. As in the first paragraph, (1x)1(1-x)^{-1} is a quadratic irrational whose Galois conjugate is (1x)1(1-x^{\dagger})^{-1}. Since the backward continued fraction expansion of xx is periodic, the minus continued fraction expansion of (1x)1(1-x)^{-1} is periodic. Proposition A.3 yields 0<(1x)1<10<(1-x^{\dagger})^{-1}<1, and so x<0x^{\dagger}<0 as required. ∎

Acknowledgments

We thank Karma Dajani and Cor Kraaikamp for fruitful discussions during their visit to Keio University. SS was supported by the JSPS KAKENHI 24K16932, Grant-in-Aid for Early-Career Scientists. HT was supported by the JSPS KAKENHI 25K21999, Grant-in-Aid for Challenging Research (Exploratory).

References

  • [1] Jon Aaronson, Random ff-expansions, Ann. Prob. 14 (1986) 1037–1057.
  • [2] Jon Aaronson and Hitoshi Nakada, Trimmed sums for non-negative, mixing stationary processes, Stochastic Processes and Their Applications 104 (2003) 173–192.
  • [3] Roy L. Adler and Leopold Flatto, The backward continued fraction map and geodesic flow, Ergodic Theory Dynam. Systems 4 (1984) 487–492.
  • [4] Romain Aimino, Matthew Nicol, and Sandro Vaienti, Annealed and quenched limit theorems for random expanding dynamical systems, Probab. Theory Relat. Fields 162 (2015) 233–274.
  • [5] Ludwig Arnold, Random Dynamical Systems, Springer Monographs in Mathematics. Springer-Verlag, Berlin, 1998.
  • [6] Wael Bahsoun, Marks Ruziboev, and Benoît Saussol, Linear response for random dynamical systems, Adv. Math. 364 (2020) 107011.
  • [7] Rufus Bowen, Periodic points and measures for Axiom A diffeomorphisms, Trans. Amer. Math. Soc. 154 (1971) 377–397.
  • [8] Rufus Bowen, Some systems with unique equilibrium states, Math. Systems Theory 8 (1974) 193–202.
  • [9] Jérôme Buzzi, Some remarks on random zeta functions, Ergodic Theory Dynam. Systems (2002) 22 1031–1040.
  • [10] Predrag Cvitanović, Periodic orbits as the skeleton of classical and quantum chaos, Physica D 51 (1991) 138–151.
  • [11] Amir Dembo and Ofer Zeitouni, Large deviations techniques and applications, Applications of Mathematics 38, Springer, second edition (1998)
  • [12] Richard S. Ellis, Entropy, large deviations, and statistical mechanics, Grundlehren der Mathematischen Wissenschaften 271 Springer (1985)
  • [13] Roberta Fabbri, Tobias Jäger, Russel Johnson, and Gerhard Keller, A Sharkovskii-type theorem for minimally forced interval maps, Topological Methods in Nonlinear Analysis Journal of the Juliusz Schauder Center 26 2005, 163–188
  • [14] Ai-Hua Fan, Thomas Jordan, Lingmin Liao, and Michał Rams, Multifractal analysis for expanding interval maps with infinitely many branches, Trans. Amer. Math. Soc. 367 (2015) 1847–1870.
  • [15] Doris Fiebig, Ulf-Rainer Fiebig, and Michiko Yuri, Pressure and equilibrium states for countable state Markov shifts, Israel J. Math. 131 (2002) 221–257.
  • [16] Évariste Galois, Analyse algébrique. Démonstration d’un théorème sur les fractions continues périodiques, Annales de mathématiques pures et appliquées. 19 (1828-29) 294–301.
  • [17] Celso Grebogi, Edward Ott, and James A. Yorke, Unstable periodic orbits and the dimensions of multifractal chaotic attractors, Phys. Rev. A 37 (1988) 1711–1725.
  • [18] Tomoki Inoue, Invariant measures for position dependent random maps with continuous random parameters, Stud. Math. 208 (2012) 11–29.
  • [19] Marius Iosifescu and Cor Kraaikamp, Metrical theory of continued fractions, Mathematics and its Applications, 547 Kluwer Academic Publishers, Dordrecht, 2002
  • [20] Johannes Jaerisch and Hiroki Takahasi, Mixed multifractal spectra of Birkhoff averages for non-uniformly expanding one-dimensional Markov maps with countably many branches, Adv. Math. 385 (2021) 107778
  • [21] Tobias Jäger and Gerhard Keller, Random minimality and continuity of invariant graphs in random dynamical systems, Trans. Amer. Math. Soc. 368(2016) 6643–6662.
  • [22] Svetlana Katok, Continued fractions, hyperbolic geometry and quadratic forms, course notes for Math 497A, summer 2001 (accessed 24th July, 2025) http://skatok.s3-website-us-east-1.amazonaws.com/pub/reu-book.pdf
  • [23] Charlene Kalle, Tom Kempton, and Evgeny Verbitskiy, The random continued fraction transformation, Nonlinearity 30 (2017) 1182–1203
  • [24] Charlene Kalle, Valentine Matache, Masato Tsujii, and Evgeny Verbitskiy, Invariant densities for random continued fractions, J. Math. Anal. Appl. 512 (2022) 126163
  • [25] Yuri Kifer, Random ff-expansions, in: Proceedings of Symposia in Pure Mathematics, 2000.
  • [26] Cor Kraaikamp, A new class of continued fraction expansions, Acta Arith. 57 (1991), 1–39.
  • [27] R. Daniel Mauldin and Mariusz Urbański, Graph directed Markov systems. Geometry and Dynamics of Limit Sets, Cambridge Tracts in Mathematics 148 Cambridge University Press (2003)
  • [28] Yuto Nakajima and Hiroki Takahasi, Hausdorff dimension of sets with restricted, slowly growing partial quotients in semi-regular continued fractions, J. Math. Soc. Japan 77 (2025) 903–916.
  • [29] Oskar Perron, Die Lehre von den Kettenbrüchen, Second edition. Chelsea Publishing Co., New York 1950.
  • [30] Yakov Pesin and Samuel Senti, Equilibrium measures for maps with inducing schemes, J. Mod. Dyn. 3 (2008) 397–430.
  • [31] Christopher G. Pinner, More on inhomogeneous Diophantine approximation, J. Théor. Nombres Bordeaux 13 539–557 (2001)
  • [32] Henri Poincaré, Les méthodes nouvelles de la méchanique céleste, Les Grandes Classiques Gauthier-Villars, 1892.
  • [33] David Ruelle, An extension of the theory of Fredholm determinants, Publ. Math. IHÉS 72 (1990) 175–193.
  • [34] Omri Sarig, Thermodynamic formalism for countable Markov shifts, Ergodic Theory Dynam. Systems 19 (1999) 1565–1593.
  • [35] Omri Sarig, Existence of Gibbs measures for countable Markov shifts, Proc. Amer. Math. Soc. 131 (2003) 1751–1758.
  • [36] Daniel W. Stroock, Probability theory, an analytic view. Third edition, Cambridge University Press, Cambridge, 2025.
  • [37] Shintaro Suzuki and Hiroki Takahasi, Distribution of cycles for one-dimensional random dynamical systems, J. Math. Anal. Appl. 527 (2023) 127465.
  • [38] Hiroki Takahasi, Large deviation principles for countable Markov shifts, Trans. Amer. Math. Soc. 372 (2019) 7831–7855.
  • [39] Hiroki Takahasi, Large Deviation Principle for arithmetic functions in continued fraction expansion, Monatshefte für Mathematik 190 (2019) 137–152.
  • [40] Hiroki Takahasi, Uniqueness of minimizer for countable Markov shifts and equidistribution of periodic points, J. Stat. Phys. 181 (2020) 2415–2431.
  • [41] Hiroki Takahasi, Large deviation principle for the backward continued fraction expansion, Stochastic Processes and Their Applications, 144 (2022) 153–172.
  • [42] Hiroki Takahasi, Level-2 large deviation principle for countable Markov shifts without Gibbs states, J. Stat. Phys. 190 (2023) 120.
  • [43] Heinrich Tietze, Über Kriterien für Konvergenz und Irrationalität unendlicher Kettenbrüche, Math. Ann. 70 (1911) 236–265.